Download Hindbrain noradrenergic A2 neurons: diverse roles in autonomic

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Adult neurogenesis wikipedia , lookup

History of neuroimaging wikipedia , lookup

Neurotransmitter wikipedia , lookup

Neuropsychology wikipedia , lookup

Haemodynamic response wikipedia , lookup

Donald O. Hebb wikipedia , lookup

Selfish brain theory wikipedia , lookup

Holonomic brain theory wikipedia , lookup

Neurophilosophy wikipedia , lookup

Brain Rules wikipedia , lookup

Biochemistry of Alzheimer's disease wikipedia , lookup

Synaptogenesis wikipedia , lookup

Axon wikipedia , lookup

Cognitive neuroscience wikipedia , lookup

Brain wikipedia , lookup

Limbic system wikipedia , lookup

Multielectrode array wikipedia , lookup

Axon guidance wikipedia , lookup

Neuroeconomics wikipedia , lookup

Aging brain wikipedia , lookup

Single-unit recording wikipedia , lookup

Connectome wikipedia , lookup

Artificial general intelligence wikipedia , lookup

Caridoid escape reaction wikipedia , lookup

Neural oscillation wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Neuroplasticity wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Mirror neuron wikipedia , lookup

Neural coding wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Development of the nervous system wikipedia , lookup

Neural correlates of consciousness wikipedia , lookup

Sexually dimorphic nucleus wikipedia , lookup

Metastability in the brain wikipedia , lookup

Central pattern generator wikipedia , lookup

Nervous system network models wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Pre-Bötzinger complex wikipedia , lookup

Optogenetics wikipedia , lookup

Neuroanatomy wikipedia , lookup

Synaptic gating wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Circumventricular organs wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Transcript
Am J Physiol Regul Integr Comp Physiol 300: R222–R235, 2011.
First published October 20, 2010; doi:10.1152/ajpregu.00556.2010.
Review
Hindbrain noradrenergic A2 neurons: diverse roles in autonomic, endocrine,
cognitive, and behavioral functions
Linda Rinaman
Department of Neuroscience, University of Pittsburgh, Pittsburgh, Pennsylvania
Submitted 24 August 2010; accepted in final form 13 October 2010
catecholaminergic; dorsal vagal complex; nucleus of the solitary tract; stress;
visceral
THE MAMMALIAN BRAIN STEM CONTAINS several distinct groups of
noradrenergic (NA) neurons that were initially described by
Dahlström and Fuxe (59) who labeled the groups A1 through
A7 as they extend from the caudal ventrolateral medulla
through the rostral lateral pons. NA neurons collectively
project throughout the central nervous system and are broadly
implicated in behavioral and physiological processes related to
attention, arousal, motivation, learning and memory, and homeostasis. NA neurons are distinguished by positive immunolabeling for tyrosine hydroxylase (TH), the rate-limiting enzyme for dopamine synthesis, and dopamine-␤-hydroxylase
(DbH), the enzyme that converts dopamine to norepinephrine
(NE) (8). Conversely, neurons comprising the A1-A7 cell
groups are not immunopositive for phenylethanolamine Nmethyltransferase (PNMT). PNMT catalyzes the synthesis of
epinephrine from NE, and its presence is used to identify
adrenergic neurons of the C1-C3 cell groups (59).
This review focuses on the A2 cell group, a fascinating
collection of NA neurons contained within the dorsal vagal
complex (DVC) in the caudal dorsomedial medulla (see Fig. 1). As
Address for reprint requests and other correspondence: L. Rinaman, Dept. of
Neuroscience, Univ. of Pittsburgh, A210 Langley Hall, Pittsburgh, PA 15260
(e-mail: [email protected]).
R222
discussed further, below, the intra-DVC location of A2 neurons
supports their known involvement in vagal sensory-motor
reflex arcs and vagal motor outflow to multiple visceral targets.
Perhaps less well appreciated is the role of A2 neurons in
processes as diverse as satiation, sickness behavior, affective
state, endocrine and behavioral stress responses, immune-tobrain signaling, emotional learning, memory consolidation,
and addictive drug dependence. A2 neurons participate in
reciprocal connections between the visceral DVC and other
medullary, pontine, diencephalic, and telencephalic brain regions that underlie these diverse processes. Direct projections
from the cortex, limbic forebrain, and hypothalamus to the
region of the A2 cell group provide a route through which
emotional and cognitive events can modulate visceral responses to diverse threats and opportunities to which the
organism is exposed, including conditioned responses that are
based on past experience (193, 225). In turn, ascending projections from A2 neurons provide a route through which
interoceptive feedback from the body impacts not only hypothalamic functions but also emotional and cognitive processing
(21, 141, 211, 225).
The neuroanatomical and phenotypic features of A2 neurons
will first be considered in this review, followed by a summary
of evidence that A2 neurons provide a critical brain-body
interface linking emotional/cognitive events with physiological
0363-6119/11 Copyright © 2011 the American Physiological Society
http://www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
Rinaman L. Hindbrain noradrenergic A2 neurons: diverse roles in autonomic, endocrine, cognitive, and behavioral functions. Am J Physiol Regul
Integr Comp Physiol 300: R222–R235, 2011. First published October 20, 2010;
doi:10.1152/ajpregu.00556.2010.—Central noradrenergic (NA) signaling is
broadly implicated in behavioral and physiological processes related to attention,
arousal, motivation, learning and memory, and homeostasis. This review focuses on
the A2 cell group of NA neurons, located within the hindbrain dorsal vagal complex
(DVC). The intra-DVC location of A2 neurons supports their role in vagal
sensory-motor reflex arcs and visceral motor outflow. A2 neurons also are reciprocally connected with multiple brain stem, hypothalamic, and limbic forebrain
regions. The extra-DVC connections of A2 neurons provide a route through which
emotional and cognitive events can modulate visceral motor outflow and also a
route through which interoceptive feedback from the body can impact hypothalamic functions as well as emotional and cognitive processing. This review
considers some of the hallmark anatomical and chemical features of A2 neurons,
followed by presentation of evidence supporting a role for A2 neurons in modulating food intake, affective behavior, behavioral and physiological stress responses, emotional learning, and drug dependence. Increased knowledge about the
organization and function of the A2 cell group and the neural circuits in which A2
neurons participate should contribute to a better understanding of how the brain
orchestrates adaptive responses to the various threats and opportunities of life and
should further reveal the central underpinnings of stress-related physiological and
emotional dysregulation.
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
distinguish these levels from the more rostral gustatory NST
(143). The visceral NST is a key component of the DVC,
which also includes the area postrema (AP) and dorsal motor
nucleus of the vagus. The DVC is a critical central node for
controlling hormonal and autonomic outflow, and relaying
interoceptive feedback from body to brain (201, 205, 206,
291). The AP and a significant portion of the medial NST
underlying the AP contain fenestrated capillaries, allowing
blood-borne factors (e.g., hormones, toxins, cytokines) to affect A2 and other neurons local to this region. Within the DVC,
AP neurons innervate the subjacent NST, and NST neurons
innervate other NST neurons as well as vagal preganglionic
parasympathetic neurons whose cell bodies occupy the dorsal
motor nucleus of the vagus and whose dendrites ramify widely
within the NST (239). All three components of the DVC also
receive extrinsic neural inputs from the periphery and brain,
described further, below.
Although the A2 cell group is centered within the visceral
NST, A2 neurons are not confined to cytoarchitecturallydistinct NST subnuclei. Instead, they form two bilaterally
symmetrical loose linear columns of medium-sized ovoid or
multipolar cells that extend rostrocaudally through the visceral
NST (232) (Fig. 1). A2 neurons are most prevalent within the
medial subnucleus of the NST at the rostrocaudal level of the
AP, but they also exist within the NST commissural subnucleus at the level of the AP and more caudally. Furthermore,
some A2 neurons are located within and just lateral to the
cytoarchitectural boundaries of the dorsal motor nucleus of
the vagus (232). The most caudal A2 neurons are located in
the upper cervical spinal cord, and the most rostral are
located rostral to the AP at the level of the caudal fourth
ventricle (Fig. 1). It should here be noted that the more rostral
A2 neurons are intermixed with PNMT-positive adrenergic
neurons of the C2 cell group (Figs. 1 and 2), which also are
TH- and DbH-positive. Many of the neuroanatomical and
functional studies cited in this review relied on anatomical
localization together with TH or DbH immunolabeling to
support, especially during stressful events that challenge bodily
homeostasis. This general theme will be supported by briefly
reviewing the involvement of A2 neurons in food intake,
affective behavior, stress responses, emotional learning, and
drug dependence. Most of the information reviewed in this
article was derived from studies using rats and, to a lesser
extent, mice; however, central NA circuits are highly conserved across mammalian species. Thus, understanding the
functional organization of the A2 cell group in rodents has
clinical relevance, and should contribute to a better understanding of stress-related physiological and emotional dysregulation
in humans.
Anatomical and Neurochemical Features
Location of A2 neurons. The A2 cell group is centered
within intermediate and caudal levels of the nucleus of the
solitary tract (NST) (Fig. 1), referred to as the visceral NST to
AJP-Regul Integr Comp Physiol • VOL
Fig. 2. Dual immunofluorescence labeling for DbH (red) and PNMT (green) in
a tissue section through the rat dorsomedial medulla (⬃13.3 mm caudal to
bregma) where neurons of the A2 and C2 cell groups intermingle. 4, fourth
ventricle.
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
Fig. 1. Immunoperoxidase localization of dopamine-␤-hydroxylase (DbH; left)
and phenylethanolamine N-methyltransferase (PNMT; right ) within the caudal
dorsomedial medulla of an adult male Sprague-Dawley rat. Side-by-side panels
represent closely adjacent tissue sections. Top 4: caudal levels of the A2 cell
group (⬃14.6 mm caudal to bregma). Bottom 4: most rostral levels (⬃13.3 mm
caudal to bregma). DbH-positive neurons within the A2 cell group are
indicated by arrows. AP, area postrema; cc, central canal; DMV, dorsal motor
nucleus of the vagus; NST, nucleus of the solitary tract; 4, fourth ventricle.
R223
Review
R224
HINDBRAIN NORADRENERGIC A2 NEURONS
Extrinsic Inputs and Axonal Projections
Sensory inputs to A2 neurons. In addition to local axonal
inputs from the superjacent AP that are positioned to relay
blood-borne signals to A2 neurons (54, 118, 238), the A2
region of the visceral NST receives sensory feedback from the
cardiovascular, respiratory, and alimentary systems (119).
These visceral sensory inputs arrive predominantly via glutamatergic glossopharyngeal and vagal afferents whose central
axons converge in the solitary tract before synapsing with the
AJP-Regul Integr Comp Physiol • VOL
dendrites and somata of NST and vagal motor neurons (5, 14,
208, 246). In mice, ⬃90% of A2 neurons receive direct
synaptic input from visceral afferents in the solitary tract (6).
These glutamatergic inputs produce tightly synced, large-amplitude excitatory postsynaptic currents in A2 neurons, providing high-fidelity transmission of sensory afferent activity (6).
Other visceral and somatic sensory inputs are relayed to the A2
region of the NST from the spinal cord, trigeminal and related
nuclei, and reticular formation (5, 7, 71, 152, 153).
Given the diversity of sensory inputs received by A2 neurons, it is not surprising that they respond to a broad array of
interoceptive signals, including hormonal, osmotic, gastrointestinal, cardiovascular, respiratory, and inflammatory signals
(20, 23, 43, 45, 66, 77, 97, 112, 120, 167, 168, 201, 202, 207,
212, 213, 230). In these and many other studies, stimulusinduced A2 neuronal activation is characterized by immunocytochemical localization of the immediate-early gene product,
Fos, together with immunolabeling for TH or DbH. Increased
Fos immunolabeling alone cannot reveal the circuits through
which A2 neurons are recruited by a given stimulus or event,
but A2 neurons are consistently activated by treatments or
situations that present actual or anticipated threats to bodily
homeostasis. In many cases the relevant information is communicated to A2 neurons by visceral sensory afferents, but in
other cases A2 neurons appear to be recruited by descending
inputs from the hypothalamus and limbic forebrain (30, 67, 68,
137, 138). These inputs are reviewed in the following section.
Central inputs to A2 neurons. Retrograde and anterograde
tract-tracing studies have revealed a wide array of brain stem
and forebrain nuclei that project directly to the visceral NST
and may participate in recruitment of A2 neurons. As summarized in Table 1, these include various regions of the medullary, pontine, and mesencephalic reticular formation (13, 157,
166, 180); the cerebellar fastigial nucleus (180); the raphé
obscurus, pallidus, magnus, paragigantocellularis, and parapyramidal region (144, 180, 256); the laterodorsal tegmental
nucleus (180); the retrotrapezoid nucleus (220); the parabrachial nucleus and Kölliker-Fuse nucleus (130, 180); the periaqueductal gray (180); the hypothalamic tuberomammillary
nuclei (182); the hypothalamic arcuate nucleus (266, 295); the
PVN (100, 204, 228, 245, 266); the lateral hypothalamic area
(180, 266, 294); the median preoptic nucleus (180); the dorsomedial hypothalamus (287); the central nucleus of the amygdala (CeA) and anterolateral bed nucleus of the stria terminalis
(alBST) (116, 140, 180, 222, 266); the lateral septal nucleus
(180); and glutamatergic pyramidal neurons in the insular
cortex and in prelimbic and infralimbic regions of the medial
prefrontal cortex (266). All of these brain regions are logical
candidates for sources of direct input to A2 neurons, although
dual-labeling electron microscopy is necessary to confirm synaptic connectivity. The limited ultrastructural evidence that is
available indicates that at least a subset of A2 neurons receive
synaptic input from noncatecholaminergic neurons in the adjacent AP (118), and at least some A2 neurons receive glutamatergic inputs from the infralimbic region of the medial
prefrontal cortex (94). It also seems likely that A2 neurons are
among the catecholaminergic neurons within the visceral NST
that receive synaptic input from the CeA (189) and from
orexin-positive neurons in the lateral hypothalamus (17).
Axonal projections of A2 neurons. A2 neurons project locally within the DVC and medullary reticular formation, and
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
identify and/or lesion A2 neurons; however, these criteria do
not allow A2 and C2 neurons to be distinguished within
visceral NST regions where they overlap. The extent to which
the connections and functions of these rostral A2/caudal C2
neurons are similar or unique remains largely unexplored.
Transmitter coexpression by A2 neurons. When considering
the functional role of A2 projection systems, it’s important to
keep in mind that these NA neurons release more than just NE
from their axon terminals and varicosities. In addition to TH
and DbH, A2 neurons express mRNA and/or are immunopositive for many additional signaling molecules. In rats, ⬃80%
of A2 neurons reportedly express mRNA for a homolog of the
vesicular glutamate transporter-2 (248), suggesting that these
neurons release glutamate along with NE. In addition, virtually
all A2 neurons are immunoreactive for prolactin-releasing
peptide (PrRP) (52). PrRP receptors are expressed within the
paraventricular nucleus of the hypothalamus (PVN) and other
brain regions targeted by A2 neurons (284), and there is
evidence that PrRP acts synergistically with NE to activate
hypophysiotropic corticotropin-releasing hormone (CRH) neurons at the apex of the hypothalamic-pituitary-adrenal (HPA)
axis (147, 261). Interestingly, the ratios of PrRP to NA biosynthetic enzymes in A2 neurons are modulated by estrogen
and stress (235), and A2 neurons express receptors for estrogen
and glucocorticoids (56, 101, 198, 243).
Regarding other peptides and phenotypic markers, subpopulations of A2 neurons are immunopositive for neuropeptide Y
(83, 233), nesfatin-1 (23), dynorphin (40), neurotensin (200),
and/or pituitary adenylate cyclase-activating polypeptide (61).
Conversely, A2 neurons apparently do not colocalize galanin
(136), somatostatin (226, 227), enkephalin (226), inhibin-␤
(226), glucagon-like peptide 1 (133, 203), or the enzyme
11-␤-hydroxysteroid dehydrogenase-2 (99), despite their close
anatomical proximity to NST neurons, which do express these
various phenotypic markers. A2 neurons also do not appear to
coexpress cocaine and amphetamine-related transcript, neurotensin, or cholecystokinin, in contrast to coexpression of these
peptides by PNMT-positive neurons of the partially overlapping C2 cell group (84, 122).
Receptor expression by A2 neurons. Neurons within the
A2 region of the visceral NST express receptor mRNA and
binding sites for a large number of neurotransmitters and
other signaling molecules, although confirmation of receptor
expression by identified A2 neurons is relatively limited.
The available data indicate that A2 neurons express ␣-2a
adrenergic receptors (154, 221) as well as receptors for
glutamate (6, 89), GABA (116), cannabinoids (38), CRH
(171), neuropeptide Y (288), leptin (80), glucocorticoids
(219), and estrogen (56, 198, 243). A2 neurons do not
express mineralocorticoid receptors (99).
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
Table 1. Central connections of the A2 region of the dorsal
vagal complex
Central Connections
incidence of nonsynaptic NA varicosities in these regions is
much higher (15, 46, 82, 175, 187, 190). Thus, A2 neurons
may release their transmitter synaptically and in a paracrine
manner, requiring that NE and other costored transmitters must
diffuse short distances through the extracellular space to bind
to cognate receptors (cf. Ref. 190).
A subset of individual A2 neurons have axon collaterals that
innervate both the PVN as well as the CeA and/or alBST in rats
(16, 20, 185, 234). In addition, some A2 neurons project both
to brain stem autonomic regions and to limbic forebrain targets
(197). Interestingly, however, different brain stem autonomic
regions appear to be targeted by different sets of A2 neurons
(111), suggesting a higher degree of anatomical specificity for
brain stem projections vs. hypothalamic and limbic forebrain
projections of A2 neurons.
A2 axonal projections that ascend rostrally beyond the
medulla do so primarily within the ventral NA bundle (VNAB)
(4, 95, 156, 160, 231, 250, 276). It’s relevant to note here that
non-NA projections from the NST to the caudal ventrolateral
medulla (111) allow visceral signals to recruit NA neurons of
the A1 cell group (14, 111, 123, 260, 286), located at the same
rostrocaudal level as the more dorsally situated A2 cell group.
Some A1 and non-NA neurons within the ventrolateral medulla
project back to the DVC (157) to participate in vagal motor
outflow to the stomach (109, 110) and presumably other
visceral targets, but the axons of many A1 neurons join A2
projections within the VNAB (44, 230, 232). The extent to
which A1 and A2 projection targets are similar or distinct
remains ripe for investigation, although there is evidence that
they target phenotypically distinct neurons and subregions of
the PVN and supraoptic nuclei (195, 196, 232). In the absence
of specific evidence to discriminate between A1 and A2 neurons, a conservative approach dictates that projections and
functions ascribed to either cell group should be considered
Italicized ⫽ source of axonal input to the A2 region; Underlined ⫽ target of
A2 axonal projections; Bold ⴝ both a source of axonal input to the A2
region and a target of A2 axonal projections. Citations are provided in text.
comprise a subset of preautonomic NST neurons implicated in
vagal control of cardiovascular and digestive functions (77,
110, 146, 184, 218, 259). Dual-labeling retrograde tracing
studies indicate that identified A2 neurons also project to
multiple higher brain regions (201), as summarized in Table 1.
These regions include the parabrachial nucleus, locus coeruleus (LC) and peri-LC region, periaqueductal gray, ventral
tegmental area and retrorubral field, midline thalamic nuclei,
tuberomammillary nucleus, arcuate nucleus, dorsomedial nucleus of the hypothalamus, PVN (Fig. 3), lateral hypothalamic
area, median preoptic nucleus, subfornical organ, supraoptic
nucleus, CeA, alBST, substantia innominata, and nucleus accumbens (NAcc) (14, 46, 82, 96, 104, 105, 116, 121, 158,
199 –201, 232, 253, 254, 272).
Regarding the central targets of A2 axonal projections, it is
useful to know that synaptic junctions may not be the primary
release site for NE and other signaling molecules that are
synthesized by A2 neurons. TH- and DbH-positive NA terminals within the PVN and other brain regions targeted by A2
neurons have been observed to form classical type I and type
II synaptic inputs to postsynaptic structures; however, the
AJP-Regul Integr Comp Physiol • VOL
Fig. 3. Dual immunofluorescence labeling of anterogradely-transported
Phaseolus vulgaris leucoagglutinin (PhAL) neural tracer (green) and DbH
(red) identifying axons in the medial parvocellular (mp) and lateral magnocellular (lm) subregions of the paraventricular nucleus of the hypothalamus
(PVN). PhAL was microinjected iontophoretically into a portion of the A2
region in an adult male Sprague-Dawley rat 14 days before death (see Ref.
201).
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
• Spinal cord, cranial nerves
Dorsal horn and lamina X
Glossopharyngeal and vagal sensory afferents (via solitary tract)
• Medulla, Pons, and Midbrain
Dorsal motor nucleus of the vagus
Nucleus ambiguous
Area postrema
Trigeminal and related nuclei
Medullary, pontine, mesencephalic reticular formation
Raphé obscurus, pallidus, magnus, paragigantocellularis, and
parapyramidal region
Locus coeruleus and peri-locus coeruleus region
Cerebellar fastigial nucleus
Parabrachial nucleus
Kölliker-Fuse nucleus
Laterodorsal tegmental nucleus
Ventral tegmental area
Retrorubral field
Retrotrapezoid nucleus
Periaqueductal gray
• Thalamus and hypothalamus
Midline thalamic nuclei
Tuberomammillary nucleus
Arcuate nucleus
Paraventricular nucleus
Lateral hypothalamic area
Dorsomedial nucleus
Median preoptic nucleus
Supraoptic nucleus
Subfornical organ
• Telencephalon
Central nucleus of the amygdala
Anterolateral bed nucleus of the stria terminalis
Substantia innominata
Nucleus accumbens
Lateral septal nucleus
Insular cortex
Medial prefrontal cortex
R225
Review
R226
HINDBRAIN NORADRENERGIC A2 NEURONS
likely shared by the other. The following sections review
several examples of functions in which A2 neurons have been
implicated, but the reader should consider that A1 neurons also
are likely to be involved in at least a subset of these functions.
A2 Neurons and Food Intake
Fig. 4. Dual immunoperoxidase labeling of cFos protein (blue/black nuclear
label) and cytoplasmic DbH (brown label) within the caudal visceral NST. This
section was taken from a rat perfused with fixative 60 min after intraperitoneal
administration of cholecystokinin octapeptide (100 ␮g/kg body wt) to stimulate vagal sensory inputs to the NST. Arrows point out activated (i.e.,
cFos-positive) A2 neurons (see Refs. 210 and 213).
AJP-Regul Integr Comp Physiol • VOL
The trajectory and targets of NA fibers within the VNAB are
distinct from those of the dorsal NA bundle, which originates
from neurons within the pontine LC (A4 and A6 cell group
regions). The vast majority of publications considering the role
of central NA signaling in stress, cognition, and affective
processes emphasize the LC and its projection targets, and
either disregard or downplay the contributions of A2 neurons
and their projections. The LC is assumed to be the principle
source of the central NA signaling that underlies not only
behavioral arousal but also HPA axis hyperactivity associated
with stress (48 –50, 289) as well as the dysregulated NA
transmission that contributes to diverse models of stress vulnerability and affective disorders (9, 51). However, NA inputs
to the PVN, CeA, alBST, and NAcc that are critical for
hormonal, behavioral, and affective responses to physiologically significant events arise from the caudal medullary A1 and
A2 cell groups, with relatively little input from the LC (70,
230, 232). When clinical researchers measure growth hormone
responses to an adrenergic agent (e.g., clonidine) to indirectly
assess brain NA signaling in individuals with stress-related
affective disorders (1, 72, 241, 242), it is primarily NA signaling by medullary A1/A2 inputs to the hypothalamus that is
being assessed (47, 81, 165, 282), not LC inputs. Furthermore,
results from a recent retrograde tracing and VNAB lesion study
indicate that the majority of NA inputs to reward-related
midbrain dopamine neurons arise from the caudal medulla,
including the A2 cell group (151). Conversely, A2 neurons do
not innervate most of the brain regions that are innervated by
the LC, including the olfactory bulb, cerebral cortex, hippocampus, medial BST, basolateral amygdala, most thalamic
nuclei, and the cerebellum (163). Moreover, NA signaling
within LC-innervated brain regions can produce behavioral
effects that are quite different from those produced by NA
signaling within A2-innervated regions. For example, ␣-adrenergic receptor activation underlies positively motivated exploratory/approach behavior in cortical and subcortical regions
innervated by the LC, but underlies stress reactions with
behavioral inhibition in regions innervated by the A2 cell
group (29, 41, 42, 69, 164, 192, 215, 247, 277, 278).
Neurons within the A2 region of the visceral NST project to
the LC (201), but project more densely to the peri-LC region
where the dendrites of LC neurons cluster, synapsing there on
TH-positive dendrites (264, 265). As mentioned previously,
A2 neurons coexpress PrRP immunolabeling (52), and the LC
contains PrRP-positive fibers and terminals as well as the
receptor (UHR-1) for PrRP (284). While the evidence is
indirect, these findings indicate that A2 neurons may provide
modulatory control over NA neurons within the LC, thereby
indirectly modifying NA signaling within cortical, hippocampal, amygdalar, thalamic, and cerebellar targets of the LC.
Interestingly, despite the LC-centered focus of most research
on the role of NA brain systems in affective behavior, researchers have not been able to demonstrate unequivocally the
necessity of LC neurons in fear, anxiety, or depressive-like
behavior (114). Indeed, LC lesions appear to increase, rather
than decrease, novelty-induced fear and anxiety in rats (103,
148). Moreover, in one study, LC lesions increased the antidepressant-like effect of reboxetine (a NE reuptake inhibitor),
whereas VNAB lesions abolished the drug’s antidepressant
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
Central NA signaling pathways, including those that arise
from A2 neurons, appear to be essential for inhibiting or
stimulating food intake under different conditions (2, 62, 142,
156, 170, 203, 209, 214, 216). It seems likely that different
subpopulations of A2 neurons with distinct axonal projections
are recruited by signals that increase or decrease food intake,
perhaps because different subpopulations target different brain
regions, and/or because different combinations of adrenergic
receptors are expressed within those regions (277). Interestingly, A2 neurons in rats appear to be activated in every
experimental situation in which food intake is inhibited, including normal satiety (37, 115, 203, 207, 262). A2 neurons are
recruited in a graded manner in rats after voluntary food intake,
such that larger meals activate larger numbers of A2 neurons
(207). Not only are A2 neurons robustly activated in rats after
systemic administration of cholecystokinin octapeptide (210,
213) (Fig. 4), they are necessary for the ability of cholecystokinin to inhibit food intake and to activate neurons within the
hypothalamic PVN (203). A2 neurons also contribute importantly to the hypophagic effect of lithium chloride (209). Food
intake is reduced in rats after central administration of PrRP
(134) or nesfatin-1 (174), each of which is coexpressed by A2
neurons (23, 52). A2 neurons are robustly activated by experimental treatments or situations that produce hypophagia or
anorexia as a part of the depressive-like sickness behavior
accompanying systemic infection or visceral malaise (23, 96 –
98, 201, 202). These conditions also are associated with inhibition of vagally mediated gastric emptying, which likely
underlies or contributes to hypophagia (37, 202, 203). The
potential involvement of A2 neurons in other aspects of food
intake and regulation of body energy homeostasis are the
subject of a recent review (201).
A2 Neurons, the LC, and Affective Behavior
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
effects (53). These results invite a continued expansion of
research into the role of A2 neurons and their central projections in affective behavior.
A2 Neurons and Stress Responses
AJP-Regul Integr Comp Physiol • VOL
mpPVN markedly attenuate CRH neuronal responses to interoceptive signals (20, 90, 137, 203, 209, 215, 234). Conversely,
chronic stress sensitizes HPA axis responses to central NA
(183) and increases the density of glutamatergic and NA
synaptic inputs to CRH-positive neurons, evidence for enhanced signaling capacity (86). The authors of the latter study
did not consider whether the increased glutamatergic and NA
inputs arise from the same A2 neurons, but this seems likely.
Magnocellular oxytocin neurons and parvocellular thyrotropin releasing hormone-positive mpPVN neurons also receive synaptic input from the A2 cell group (57, 58, 93, 240,
283). In addition, nonendocrine gastric preautonomic neurons
in the PVN receive direct synaptic input from NA nerve
terminals that include inputs from A2 neurons (15). Preautonomic PVN neurons project to hindbrain and spinal centers to
control autonomic motor outflow to the gastrointestinal tract
and other organ systems (16, 87, 204, 251) to thereby shape
visceral responses to emotive stimuli and stress (228, 292).
alBST. The anterolateral group of BST nuclei (alBST) includes the juxtacapsular, oval, rhomboid, fusiform, and subcommissural zone (75). The alBST is connected with autonomicrelated portions of the hypothalamus and caudal medulla, and
receives an extremely dense NA innervation that arises from
A2 (and A1) neurons, but not from the LC (11, 25, 69, 74, 75,
186 –188, 252, 275). NA acts within the alBST to modulate
behavioral, hormonal, and conditioned emotional responses to
stress (41, 179), including, for example, responses to the stress
of precipitated opiate withdrawal (11, 78). The alBST also
receives input from the hippocampus and prefrontal cortex, and
has abundant projections to the mpPVN (74). At least a subset
of A2 neurons that innervate the alBST have axon collaterals
that target the mpPVN (16, 20), and NA signaling within the
alBST contributes to stress-induced HPA axis activation (88,
135). Certain aspects of BST-mediated anxiety responses appear to depend on CRH inputs from the amygdala (63, 270,
271) with which the alBST is strongly and reciprocally connected (73). Blockade of NA signaling in the ventral alBST
reduces immobilization stress-induced anxiety in the elevatedplus maze and attenuates immobilization stress-induced increases in plasma ACTH, but the same pharmacological manipulation has no effect to attenuate stress responses in a
subsequent social interaction test (41); the reverse is true for
similar manipulations in the CeA (42). These findings suggest
that A2 inputs to the alBST and CeA are involved in different
specific components of stress and anxiety responses (cf. Ref.
244).
CeA. The CeA, like the alBST, is a subcortical limbic
structure characterized by its extensive connections with the
hypothalamus and with brain stem viscerosensory and autonomic control nuclei. As part of the striatal amygdala, CeA
neurons are primarily GABAergic and coexpress CRH, similar
to neurons in the alBST (73, 249). Although NA inputs to the
CeA are significantly less dense than NA inputs to the mpPVN
and alBST (16, 167, 169), NA signaling within the CeA
modulates behavioral responses to stressful events, including
fear, anxiety, and avoidance behavior (64, 125). NA signaling
in the CeA increases during precipitated opiate withdrawal and
contributes to the negative affective (i.e., aversive) consequences of withdrawal (273), similar to increased NA signaling
and implication in aversive effects within the alBST (11). A2
neurons that project to the CeA are activated in rats after
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
Stressors are stimuli or events that challenge (or are perceived to challenge) bodily homeostasis and well-being. Signals generated by stressors may initially arrive from the environment (e.g., visual or olfactory signals), or may arise from
within the body (e.g., cardiovascular or gastrointestinal signals). Physiological and behavioral responses to stressful stimuli are the product of interactions among multiple brain regions
(107, 108). However, three interconnected regions of the hypothalamus and so-called extended amygdala are especially
important, and have been the subject of extensive experimental
attention: the medial parvocellular PVN (mpPVN), CeA, and
alBST. Each region contains CRH neurons that receive synaptic input from NA terminals arising primarily or exclusively
from the A2 (and A1) cell groups, and CRH neuronal activity
within each region is closely regulated by these inputs (4, 10,
65, 74, 78, 79, 124, 126 –128, 139, 181, 186, 191, 192, 244).
A2 neurons express glucocorticoid receptors (101), and central
adrenergic and CRH receptors are regulated by glucocorticoids, which are known to affect NE synthesis and turnover
throughout the brain. NA-CRH signaling pathways are viewed
as part of a central adrenal steroid-sensitive network that tunes
physiological and behavioral responses during conditions of
acute or chronic stress (106). In general, and as discussed
further, below, NA signaling is pivotal in facilitating HPA axis
and behavioral responses to stress, and can modulate unconditioned and conditioned behavioral responses to stressful and
emotional stimuli, including stimuli that evoke fear and anxiety
(223). Enhanced NA transmission in human subjects is associated with enhanced HPA axis responses to stress, which may
contribute to the psychopathology of depression, anxiety, and
other affective disorders (128). The hindbrain A2 cell group
appears to be a fundamental player in these central mechanisms, as summarized in the following sections.
PVN. Most hypothalamic NA input arrives from medullary
NA cell groups. A2 neurons appear to selectively target the
mpPVN (55, 230, 232), although axonal projections from the
A2 region to the lateral magnocellular PVN also are common
(see Fig. 3 and discussion in the following section). The
necessity of NA inputs for HPA axis responses to stress
appears to vary across different types of stress stimuli (20, 66,
137, 224, 229), but NA input to the mpPVN, arriving via the
VNAB, provides the major known stimulus for CRH synthesis
and release (65, 181, 192, 250, 283). CRH is the principal and
obligate hypophysiotropic peptide driving the pituitary-adrenal
axis under basal conditions and in response to homeostatic
challenge (192, 274).
In rats, stressful stimuli that activate A2 neurons and recruit
the HPA axis also activate hypothalamic oxytocin neurons
(177, 178, 206, 263, 267–269). CRH and oxytocin neurons
receive direct synaptic input from NA terminals, and NA
inputs increase CRH and oxytocin excitability (3, 4, 22, 113,
155, 192, 195, 196, 217, 285). In late pregnancy and during
lactation, oxytocin and HPA axis responses to stressors are
attenuated by mechanisms that reduce NA tone within the PVN
(26 –28, 76, 257, 258). Lesions that decrease NA input to the
R227
Review
R228
HINDBRAIN NORADRENERGIC A2 NEURONS
gastric vagal sensory stimulation with exogenous cholecystokinin (169) or systemic immune challenge (97), and also are
activated by emotionally salient exteroceptive stimuli, such as
exposure to a predator odor (168).
A2 Neurons and Emotional Learning
A2 Neurons in Drug Use and Dependence
Central neural adaptations elicited by exposure to addictive
drugs are not limited to brain reward circuits, but also are
manifest in stress-related pathways that are implicated in addiction (173). For example, rats dependent on morphine display increased enzymatic activity of A2 neurons and increased
NA turnover within the PVN, concurrent with enhanced activity of the HPA axis that depends on NA input (18, 92, 131, 132,
172). In addition, medullary NA inputs to the alBST, CeA, and
NAcc are critical for the aversiveness of acute opiate withdrawal, and for stress-induced relapse of drug seeking for
opiates, cocaine, ethanol, and nicotine (11, 24, 36, 69, 78, 145,
244, 293). NA inputs to the alBST trigger GABAergic inhibition of alBST neurons that project to the ventral tegmental
area, which likely contributes to the inhibition of DA neurons
that occurs during opiate withdrawal (78). Thus, common
inputs to the hypothalamus and limbic forebrain from the A2
cell group could be a critical factor linking these brain areas in
circuits that underlie drug use and dependence (102, 236, 237).
Even 5 wk after opiate withdrawal, neurons within these limbic
forebrain regions remain hypersensitive to drug-related cues
and stress, which may drive behavior away from the pursuit of
AJP-Regul Integr Comp Physiol • VOL
Conclusion
The diverse challenges and opportunities of life elicit a
constellation of autonomic, endocrine, cognitive, and behavioral responses, and A2 neurons are poised to contribute to the
central coordination and modulation of these responses. As
reviewed in this article, feedback regarding the body’s physiological state is relayed by A2 neurons to multiple regions of
the brain stem, hypothalamus, and limbic forebrain. The A2
cell group is best defined by its afferent and efferent connections within a complex neural network that extends from the
spinal cord to the cortex. The available evidence indicates that
by virtue of their central axonal projections, this relatively
small group of caudal medullary neurons can modulate ongoing and future physiological processes and behaviors, and may
also contribute to the affective, contextual, and cognitive attributes of experience that depend on interoceptive feedback
from body to brain (19, 60, 149, 194, 255).
GRANTS
This manuscript was prepared with the support of National Institute of
Mental Health Grants MH-59911 and MH-081817.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the author(s).
REFERENCES
1. Abelson JL, Cameron OG. Adrenergic dysfunction in anxiety disorders. In: Adrenergic Dysfunction and Psychobiology, edited by Cameron
OG. Washington, DC: American Psychiatric Press, 1994.
2. Ahlskog JE, Hoebel BG. Overeating and obesity from damage to a
noradrenergic system in the brain. Science 182: 166 –169, 1973.
3. Al-Damluji S. Adrenergic mechanisms in the control of corticotropin
secretion. J Endocrinol 119: 5–14, 1988.
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
The effectiveness with which emotionally significant experiences are encoded into long-term memory is dependent, at
least in part, on interoceptive feedback from body to brain, and
increased NA signaling within the limbic forebrain is strongly
implicated in emotional learning (31–35, 85, 161, 162, 279,
280). As reviewed earlier in this article, interoceptive signals
funneled through the hindbrain DVC engage A2 neurons,
including those that project to the amygdala and NAcc. Both
limbic regions play a crucial role in the encoding, storage, and
retrieval of memories associated with emotionally significant
events. Activation of NA receptors within the amygdala and
NAcc influences synaptic changes that are necessary at the
time of encoding to facilitate long-term memory for emotional
events (85, 124). NE release in the amygdala has been established as a neural substrate for memory modulation elicited by
peripheral arousal (85, 280, 281), and NA inputs to the NAcc
contribute importantly to the processing of appetitive and
aversive reinforcement signals that impact learning and memory (70). Neurons within the amygdala and NAcc respond to
gastric and cardiovascular vagal sensory signals that engage
the A2 cell group (126, 150, 159). These A2 inputs play a key
role in modulating amygdalar and NAcc activity by releasing
NA in response to heightened states of arousal, providing a
clear anatomical route through which emotional (i.e., visceral)
signals can modulate learning and memory (124).
A2 inputs to other brain regions may affect learning and
memory in a more indirect manner. For example, A2 inputs to
NA neurons in the LC (264, 265) and to dopamine neurons in
the midbrain retrorubral field (151) that innervate the entorhinal cortex and hippocampus may contribute to the modulation
of declarative and spatial memory processes.
natural rewards, such as food and sex, and toward drug-related
rewards to thereby perpetuate a cycle of drug addiction (102).
Clinical observations of former opiate addicts revealed a prolonged hyper-responsiveness to stress, including altered cortisol release (129), which may be at least partly due to altered
function of A2 signaling pathways. For example, A2-to-NAcc
projection neurons are similarly activated by noxious visceral
stimuli and by precipitated opiate withdrawal (69, 117), and
the same projection pathway is implicated in cannabinoid
modulation of NAcc activity and cannabinoid-induced aversion (38, 39).
Increased NA release within the extended amygdala continues to influence stress and anxiety systems in the brain for
some time following acute drug withdrawal, even after somatic
signs dissipate (91). Interestingly, conditioned preference for
morphine is absent in DbH knock-out mice in which NE
synthesis is interrupted, but preference is restored if DbH is
rescued to restore NA signaling within and from the NTS, but
not the LC (176). Repeated nicotine self-administration increases NA receptor sensitivity in the PVN and also enhances
HPA function (290). Collectively, it seems that NA signaling
from the A2 cell group to the hypothalamus and limbic forebrain contributes to mechanisms that support drug seeking and
self-administration, increased anxiety during drug abstinence,
altered reward processing (i.e., dysphoria), and the general
relationship between the use of mood-altering drugs and mood
disorders (12, 38, 78, 131, 176, 244).
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
AJP-Regul Integr Comp Physiol • VOL
23. Bonnet MS, Pecchi E, Trouslard J, Jean A, Dallaporta M, Troadec
JD. Central nesfatin-1-expressing neurons are sensitive to peripheral
inflammatory stimulus. J Neuroinflammation 6: 27, 2009.
24. Brown ZJ, Tribe E, D’souza NA, Erb S. Interaction between noradrenaline and corticotrophin-releasing factor in the reinstatement of cocaine
seeking in the rat. Psychopharmacologia 203: 121–130, 2009.
25. Brownstein MJ, Palkovits M. Catecholamines, serotonin, acetylcholine,
and ␥-aminobutyric acid in the rat brain: biochemical studies. In: Handbook of Chemical Neuroanatomy, edited by Bjorklund A and Hokfelt T.
Amsterdam: Elsevier, 1984, p. 23–54.
26. Brunton PJ, McKay AJ, Ochedalski T, Piastowska A, Rebas E,
Lachowicz A, Russell JA. Central opioid inhibition of neuroendocrine
stress responses in pregnancy in the rat is induced by the neurosteroid
allopregnanolone. J Neurosci 29: 6449 –6460, 2009.
27. Brunton PJ, Russell JA. Attenuated hypothalamo-pituitary-adrenal axis
responses to immune challenge during pregnancy: the neurosteroidopioid connection. J Physiol 586: 369 –375, 2008.
28. Brunton PJ, Russell JA. Keeping oxytocin neurons under control during
stress in pregnancy. Prog Brain Res 170: 365–377, 2008.
29. Buller K, Xu Y, Dayas C, Day T. Dorsal and ventral medullary
catecholamine cell groups contribute differentially to systemic interleukin-1␤-induced hypothalamic pituitary adrenal axis responses. Neuroendocrinology 73: 129 –138, 2001.
30. Buller KM, Dayas CV, Day TA. Descending pathways from the
paraventricular nucleus contribute to the recruitment of brainstem nuclei
following a systemic immune challenge. Neuroscience 118: 189 –203,
2003.
31. Cahill L, Babinsky R, Markowitsch HJ, McGaugh JL. The amygdala
and emotional memory. Nature 377: 295–296, 1995.
32. Cahill L, McGaugh JL. Amygdaloid complex lesions differentially
affect retention of tasks using appetitive and aversive reinforcement.
Behav Neurosci 104: 532–543, 1990.
33. Cahill L, McGaugh JL. Mechanisms of emotional arousal and lasting
declarative memory. Trends Neurosci 21: 294 –299, 1998.
34. Cahill L, McGaugh JL. A novel demonstration of enhanced memory
associated with emotional arousal. Conscious Cogn 4: 410 –421, 1995.
35. Cahill L, Prins B, Weber M, McGaugh JL. ␤-Adrenergic activation
and memory for emotional events. Nature 371: 702–704, 1994.
36. Caillé S, Espejo EF, Reneric JP, Cador M, Koob GF, Stinus L. Total
neurochemical lesion of noradrenergic neurons of the locus ceruleus does
not alter either naloxone-precipitated or spontaneous opiate withdrawal
nor does it influence ability of clonidine to reverse opiate withdrawal. J
Pharmacol Exp 290: 881–892, 1999.
37. Callahan JB, Rinaman L. The postnatal emergence of dehydration
anorexia in rats is temporally associated with the emergence of dehydration-induced inhibition of gastric emptying. Physiol Behav 64: 683–687,
1998.
38. Carvalho AF, Mackie K, Van Bockstaele EJ. Cannabinoid modulation
of limbic forebrain noradrenergic circuitry. Eur J Neurosci 31: 286 –301,
2010.
39. Carvalho AF, Reyes ARS, Sterling RC, Unterwald E, Van Bockstaele
EJ. Contribution of limbic norepinephrine to cannabinoid-induced aversion. Psychopharmacology (Berl) 211: 479 –491, 2010.
40. Ceccatelli S, Seroogyb KB, Millhornc DE, Tereniusd L. Presence of a
dynorphin-like peptide in a restricted subpopulation of catecholaminergic
neurons in rat nucleus tractus solitarii. Brain Res 589: 225–230, 1992.
41. Cecchi M, Khoshbouei H, Javors M, Morilak DA. Modulatory effects
of norepinephrine in the lateral bed nucleus of the stria terminalis on
behavioral and neuroendocrine responses to acute stress. Neuroscience
112: 13–21, 2002.
42. Cecchi M, Khoshbouei H, Morilak DA. Modulatory effects of norepinephrine, acting on ␣1 receptors in the central nucleus of the amygdala,
on behavioral and neuroendocrine responses to acute immobilization
stress. Neuropharmacology 43: 1139 –1147, 2002.
43. Chan RK, Sawchenko PE. Spatially and temporally differentiated
patterns of c-fos expression in brainstem catecholaminergic cell groups
induced by cardiovascular challenges in the rat. J Comp Neurol 348:
433–460, 1994.
44. Chan RKW, Peto CA, Sawchenko PE. A1 catecholamine cell group:
fine structure and synaptic input from the nucleus of the solitary tract. J
Comp Neurol 351: 62–80, 1995.
45. Chan RKW, Sawchenko PE. Organization and transmitter specificity of
medullary neurons activated by sustained hypertension: implications for
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
4. Alonso G, Szafarczyk A, Balmefrezol M, Assenmacher I. Immunocytochemical evidence of stimulatory control by the ventral noradrenergic bundle of parvicellular neurons of the paraventricular nucleus secreting corticotropin-releasing hormone and vasopressin in rats. Brain Res
397: 297–307, 1986.
5. Altschuler SM, Bao X, Bieger D, Hopkins DA, Miselis RR. Viscerotopic representation of the upper alimentary tract in the rat: sensory
ganglia and nuclei of the solitary and spinal trigeminal tracts. J Comp
Neurol 283: 248 –268, 1989.
6. Appleyard SM, Marks D, Kobayashi K, Okano H, Low MJ, Andresen MC. Visceral afferents directly activate catecholamine neurons in the
solitary tract nucleus. J Neurosci 27: 13292–13302, 2007.
7. Arbab MA, Delgado T, Wiklund L, Svendgaard NA. Brain stem
terminations of the trigeminal and upper spinal ganglia innervation of the
cerebrovascular system: WGA-HRP transganglionic study. J Cereb
Blood Flow Metab 8: 54 –63, 1988.
8. Armstrong DM, Ross CA, Pickel VP, Joh TH, Reis DJ. Distribution of
dopamine, noadrenaline and adrenaline-containing cell bodies in the rat
medulla oblongata: demonstration by immunocytochemical localization
of catecholamine biosynthetic enzymes. J Comp Neurol 211: 173–187,
1982.
9. Asakura M, Nagashima H, Fujii S, Sasuga Y, Misonoh A, Hasegawa
H, Osada K. Influences of chronic stress on central nervous systems. Jpn
J Psychopharmacol 20: 97–105, 2000.
10. Asan E, Yilmazer-Hanke DM, Eliava M, Hantsch M, Lesch KP,
Schmitt A. The corticotropin-releasing factor (CRF)-system and monoaminergic afferents in the central amygdala: investigations in different
mouse strains and comparison with the rat. Neuroscience 131: 953–967,
2005.
11. Aston-Jones G, Delfs JM, Druhan J, Zhu Y. The bed nucleus of the
stria terminalis: a target site for noradrenergic actions in opiate withdrawal. Ann NY Acad Sci 877: 486 –498, 1999.
12. Aston-Jones G, Harris GC. Brain substrates for increased drug seeing
during protracted withdrawal. Neuropsychopharmacology 47, Suppl 1:
167–179, 2004.
13. Babic T, de Oliveira CV, Ciriello J. Collateral axonal projections from
rostral ventromedial medullary nitric oxide synthase containing neurons
to brainstem autonomic sites. Brain Res 1211: 2008.
14. Bailey TW, Hermes SM, Andresen MC, Aicher AA. Cranial visceral
afferent pathways through the nucleus of the solitary tract to caudal
ventrolateral medulla or paraventricular hypothalamus: target-specific
synaptic reliability and convergence patterns. J Neurosci 26: 11893–
11902, 2006.
15. Balcita-Pedicino JJ, Rinaman L. Noradrenergic axon terminals contact
gastric pre-autonomic neurons in the paraventricular nucleus of the
hypothalamus in rats. J Comp Neurol 501: 608 –618, 2007.
16. Banihashemi L, Rinaman L. Noradrenergic inputs to the bed nucleus of
the stria teminalis and paraventricular nucleus of the hypothalamus
underlie hypothalamic-pituitary-adrenal axis but not hypophagic or conditioned avoidance responses to systemic yohimbine. J Neurosci 26:
11442–11453, 2006.
17. Barrera G, Hernandez A, Poulin JF, Laforest S, Drolet G, Morilak
DA. Galanin-mediated anxiolytic effect in rat central amygdala is not a
result of corelease from noradrenergic terminals. Synapse 59: 27–40,
2006.
18. Benavides M, Laorden ML, García-Borrón JC, Milanés MV. Regulation of tyrosine hydroxylase levels and activity and Fos expression
during opioid withdrawal in the hypothalamic PVN and medulla oblongata catecholaminergic cell groups innervating the PVN. Eur J Neurosci
17: 103–112, 2003.
19. Berntson GG, Sarter M, Cacioppo JT. Ascending visceral regulation
of cortical affective information processing. Eur J Neurosci 18: 2103–
2109, 2003.
20. Bienkowski MS, Rinaman L. Noradrenergic inputs to the paraventricular hypothalamus contribute to hypothalamic-pituitary-adrenal axis and
central Fos activation in rats after acute systemic endotoxin exposure.
Neuroscience 156: 1093–1102, 2008.
21. Blessing WW. The Lower Brainstem and Bodily Homeostasis. New
York: Oxford University Press, 1997.
22. Bojo L, Cassuto J, Nellgard P. Pain-induced inhibition of gastric
motility is mediated by adrenergic and vagal non-adrenergic reflexes in
the rat. Acta Physiol Scand 146: 377–383, 1992.
R229
Review
R230
46.
47.
48.
49.
50.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
understanding baroreceptor reflex circuitry. J Neurosci 18: 371–387,
1998.
Chang HT. Noradrenergic innervation of the substantia innominata: a
light and electron microscopic analysis of dopamine ␤-hydroxylase
immunoreactive elements in the rat. Exp Neurol 104: 101–112, 1989.
Chapman IM, Kapoor R, Willoughby JO. Endogenous catecholamines
modulate growth hormone release in the conscious rat during hypoglycaemia but not in the basal state. J Neuroendocrinol 5: 145–150, 1993.
Charney DS, Grillon C, Bremner JD. The neurobiological basis of
anxiety and fear: circuits, mechanisms, and neurochemical interactions
(part I). Neuroscientist 4: 35–44, 1998.
Charney DS, Grillon CCG, Bremner JD. The neurobiological basis of
anxiety and fear: circuits, mechanisms, and neurochemical interactions
(part II). Neuroscientist 4: 122–132, 1998.
Charney DS, Grillon CCG, Bremner JD. The neurobiological basis of
anxiety and fear: circuits, mechanisms, and neurochemical interactions
(part III). Neuroscientist 4: 122–132, 1998.
Charney DS, Heninger GR, Redmond J DE. Yohimbine induced
anxiety and increased noradrenergic function in humans: effects of
diazepam and clonidine. Life Sci 33: 19 –29, 1983.
Chen CT, Dun SL, Dun NJ, Chang JK. Prolactin-releasing peptideimmunoreactivity in A1 and A2 noradrenergic neurons of the rat medulla. Brain Res 822: 276 –279, 1999.
Cryan JF, Page ME, Lucki I. Noradrenergic lesions differentially alter
the antidepressant-like effects of reboxetine in a modified forced swim
test. Eur J Pharmacol 436: 197–205, 2002.
Cunningham ET Jr, Miselis RR, Sawchenko PE. The relationship of
efferent projections from the area postema to vagal motor and brain stem
catecholamine-containing cell groups: an axonal transport and immunohistochemical study in the rat. Neuroscience 58: 635–648, 1994.
Cunningham ET Jr, Sawchenko PE. Anatomical specificity of noradrenergic inputs to the paraventricular and supraoptic nuclei of the rat
hypothalamus. J Comp Neurol 274: 60 –76, 1988.
Curran-Rauhut MA, Petersen SL. Oestradiol-dependent and -independent modulation of tyrosine hydroxylase mRNA levels in subpopulations
of A1 and A2 neurones with oestrogen receptor (ER)␣ and ER␤ gene
expression. J Neuroendocrinol 15: 296 –303, 2003.
Daftary SS, Boudaba C, Szabo K, Tasker JG. Noradrenergic excitation of magnocellular neurons in the rat hypothalamic paraventricular
nucleus via intranuclear glutamatergic circuits. J Neurosci 18: 10619 –
10628, 1998.
Daftary SS, Boudaba C, Tasker JG. Noradrenergic regulation of
parvocellular neurons in the rat hypothalamic paraventricular nucleus.
Neuroscience 96: 743–751, 2000.
Dahlström A, Fuxe K. Evidence for the existence of monoaminecontaining neurons in the central nervous system. I. Demonstration of
monoamines in the cell bodies of brain stem neurons. Acta Physiol Scand
62, Suppl 232: 3–55, 1964.
Damasio A. The Feeling of What Happens: Body and Emotion in the
Making of Consciousness. New York: Harcourt Brace, 1999.
Das M, Vihlen CS, Legradi G. Hypothalamic and brainstem sources of
pituitary adenylate cyclase-activating polypeptide nerve fibers innervating the hypothalamic paraventricular nucleus in the rat. J Comp Neurol
500: 761–776, 2007.
Date Y, Shimbara T, Koda S, Toshinai K, Ida T, Murakami N,
Miyazato M, Kokame K, Ishizuka Ishida Y, Kageyama H, Shioda S,
Kangawa K, Nakazato M. Peripheral ghrelin transmits orexigenic
signals through the noradrenergic pathway from the hindbrain to the
hypothalamus. Cell Metab 4: 323–331, 2006.
Davis M. Are different parts of the extended amygdala involved in fear
versus anxiety? Biol Psychiatry 44: 1239 –1247, 1998.
Davis M, Rainnie D, Cassell M. Neurotransmission in the rat amygdala
related to fear and anxiety. Trends Neurosci 17: 208 –214, 1994.
Day T, Ferguson AV, Renaud L. Noradrenergic afferents facilitate the
activity of tuberoinfundibular neurons of the hypothalamic paraventricular nucleus. Neuroendocrinology 41: 17–22, 1985.
Dayas CV, Buller KM, Crane JW, Xu Y, Day TA. Stressor categorization: acute physical and psychological stressors elicit distinctive recruitment patterns in the amygdala and in medullary noradrenergic cell
groups. Eur J Neurosci 14: 1143–1152, 2001.
Dayas CV, Buller KM, Day TA. Hypothalamic paraventricular nucleus
neurons regulate medullary catecholamine cell responses to restraint
stress. J Comp Neurol 478: 22–34, 2004.
AJP-Regul Integr Comp Physiol • VOL
68. Dayas CV, Day TA. Opposing roles for medial and central amygdala in
the initiation of noradrenergic cell responses to a psychological stressor.
Eur J Neurosci 15: 1712–1718, 2001.
69. Delfs JM, Zhu Y, Druhan JP, Aston-Jones G. Noradrenaline in the
ventral forebrain is critical for opiate withdrawal-induced aversion.
Nature 403: 430 –434, 2000.
70. Delfs JM, Zhu Y, Druhan JP, Aston-Jones GS. Origin of noradrenergic afferents to the shell subregion of the nucleus accumbens: anterograde and retrograde tract-tracing studies in the rat. Brain Res 806:
127–140, 1998.
71. de Sousa Buck H, Caous CA, Lindsey CJ. Projections of the paratrigeminal nucleus to the ambiguus, rostroventrolateral and lateral reticular
nuclei, and the solitary tract. Auton Neurosci 87: 187–200, 2001.
72. Dinan TG, Barry S, Ahkion S, Chua A, Keeling PW. Assessment of
central noradrenergic functioning in irritable bowel syndrome using a
neuroendocrine challenge test. J Psychosom Res 43: 575–580, 1990.
73. Dong HW, Petrovich GD, Swanson LW. Topography of projections
from amygdala to bed nuclei of the stria terminalis. Brain Res Rev 38:
192–246, 2001.
74. Dong HW, Petrovich GD, Watts AG, Swanson LW. Basic organization of projections from the oval and fusiform nuclei of the bed nuclei of
the stria terminalis in adult rat brain. J Comp Neurol 436: 430 –455, 2001.
75. Dong HW, Swanson LW. Organization of axonal projections from the
anterolateral area of the bed nucleus of the stria terminalis. J Comp
Neurol 468: 277–298, 2004.
76. Douglas AJ. Central noradrenergic mechanisms underlying acute stress
responses of the hypothalamo-pituitary-adrenal axis: adaptations through
pregnancy and lactation. Stress 8: 5–18, 2005.
77. Duale H, Waki H, Howorth P, Kasparov S, Teschemacher AG, Paton
JFR. Restraining influence of A2 neurons in chronic control of arterial
pressure in spontaneously hypertensive rats. Cardiovasc Res 76: 184 –
193, 2007.
78. Dumont EC, Williams JT. Noradrenaline triggers GABAA inhibition of
bed nucleus of the stria terminalis neurons projecting to the ventral
tegmental area. J Neurosci 24: 8198 –8204, 2004.
79. Dunn AJ, Swiergiel AH, Palamarchouk V. Brain circuits involved in
corticotropin-releasing factor-norepinephrine interactions during stress.
Ann NY Acad Sci 1018: 25–34, 2004.
80. Ellacott KLJ, Lawrence CB, Rothwell NJ, Luckman SM. PRLreleasing peptide interacts with leptin to reduce food intake and body
weight. Endocrinology 143: 368 –374, 2002.
81. Emanuel AJ, Ritter S. Hindbrain catecholamine neurons modulate the
growth hormone but not the feeding response to ghrelin. Neuroendocrinology 151: 3237–3246, 2010.
82. Ericson H, Blomqvist A, Köhler C. Brainstem afferents to the tuberomammillary nucleus in the rat brain with special reference to monoaminergic innervation. J Comp Neurol 281: 169 –192, 1989.
83. Everitt BJ, Hökfelt T. The coexistence of neuropeptide-Y with other
peptides and amines in the central nervous system. In: Neuropeptide Y,
edited by Mutt V, Fuxe K, Hökfelt T, and Lundberg JM. New York:
Raven, 1989, p. 61–71.
84. Fekete C, Wittman G, Liposits Z, Lechan RM. Origin of cocaine- and
amphetamine-regulated transcript (CART)-immunoreactive innervation
of the hypothalamic paraventricular nucleus. J Comp Neurol 469: 340 –
350, 2004.
85. Ferry B, Roozendaal B, McGaugh JL. Role of norepinephrine in
mediating stress hormone regulation of long-term memory storage: a
critical involvement of the amygdala. Biol Psychiatry 46: 1140 –1152,
1999.
86. Flak JN, Ostrander MM, Tasker JG, Herman JP. Chronic stressinduced neurotransmitter plasticity in the PVN. J Comp Neurol 517:
156 –165, 2009.
87. Flanagan LM, Verbalis JG, Stricker EM. Effects of anorexigenic
treatments on gastric motility in rats. Am J Physiol Regul Integr Comp
Physiol 256: R955–R961, 1989.
88. Forray MI, Gysling K. Role of noradrenergic projections to the bed
nucleus of the stria terminalis in the regulation of the hypothalamicpituitary-adrenal axis. Brain Res Rev 47: 145–160, 2004.
89. Forray MI, Gysling K, Andres ME, Bustos G, Araneda S. Medullary
noradrenergic neurons projecting to the bed nucleus of the stria terminalis
express mRNA for the NMDA-NR1 receptor. Brain Res Bull 52: 163–
169, 2000.
90. Fraley GS, Ritter S. Immunolesion of norepinephrine and epinephrine
afferents to medial hypothalamus alters basal and 2-deoxy-D-glucose300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
51.
HINDBRAIN NORADRENERGIC A2 NEURONS
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
91.
92.
93.
94.
95.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
AJP-Regul Integr Comp Physiol • VOL
110. Hermann GE, Nasse JS, Rogers RC. ␣-1 Adrenergic input to solitary
nucleus neurones: calcium oscillations, excitation and gastric reflex
control. J Physiol 562: 553–568, 2005.
111. Hermes SM, Mitchell JL, Aicher SA. Most neurons in the nucleus
tractus solitarii do not send collateral projections to multiple autonomic
targets in the rat brain. Exp Neurol 198: 539 –551, 2006.
112. Hollis JH, Lightman SL, Lowry CA. Integration of systematic and
visceral sensory information by medullary catecholaminergic systems
during peripheral inflammation. Ann NY Acad Sci 1018: 71–75, 2004.
113. Horie S, Shioda S, Nakai Y. Catecholaminergic innervation of oxytocin
neurons in the paraventricular nucleus of the rat hypothalamus as revealed by double-labeling immunoelectron microscopy. Acta Anat
(Basel) 147: 184 –192, 1993.
114. Itoi K, Sugimoto N. The brainstem noradrenergic systems in stress,
anxiety and depression. J Neuroendocrinol 22: 355–361, 2010.
115. Jelsing J, Galzin AM, Guillot E, Pruniaux MP, Larsen PJ, Vrang N.
Localization and phenotypic characterization of brainstem neurons activated by rimonabant and WIN55,212–2. Brain Res Bull 78: 202–210,
2009.
116. Jia HG, Rao ZR, Shi JW. Evidence of g-aminobutyric acidergic control
over the catecholaminergic projection from the medulla oblongata to the
central nucleus of the amygdala. J Comp Neurol 381: 262–281, 1997.
117. Jin GR, Rao ZR, Shi JW. Visceral noxious stimulation induced expression of Fos protein in medullary catecholaminergic neurons projecting to
nucleus accumbens in the rat: a study with triple labelig method of HRP
tracing combined with Fos and TH immunohistochemistry. Brain Res
648: 196 –202, 1985.
118. Kachidian P, Pickel VM. Localization of tyrosine hydroxylase in
neuronal targets and efferents of the area postrema in the nucleus tractus
solitarii of the rat. J Comp Neurol 329: 337–353, 1993.
119. Kalia M, Sullivan JM. Brainstem projections of sensory and motor
components of the vagus nerve in the rat. J Comp Neurol 211: 248 –265,
1982.
120. Kasparov S, Teschemacher AG. Altered central catecholaminergic
transmission and cardiovascular disease. Exp Physiol 93: 725–740, 2008.
121. Kawai Y, Takagi H, Yanai K, Tohyama M. Adrenergic projection
from the caudal part of the nucleus of the tractus solitarius to the
parabrachial nucleus in the rat: immunocytochemical study combined
with a retrograde tracing method. Brain Res 459: 369 –372, 1988.
122. Kawaia Y, Takagi H, Tohyama M. Co-localization of neurotensin- and
cholecystokinin-like immunoreactivities in catecholamine neurons in the
rat dorsomedial medulla. Neuroscience 24: 227–236, 1988.
123. Kawano H, Masuko S. Neurons in the caudal ventrolateral medulla
projecting to the paraventricular hypothalamic nucleus receive synaptic
inputs from the nucleus of the solitary tract: a light and electron
microscopic double-labeling study in the rat. Neurosci Lett 218: 33–36,
1996.
124. Kerfoot EC, Chattillion EA, Williams CL. Functional interactions
between the nucleus tractus solitarius (NTS) and nucleus accumbens
shell in modulating memory for arousing experiences. Neurobiol Learn
Mem 89: 47–60, 2008.
125. Khoshbouei H, Cecchi M, Dove S, Javors M, Morilak DA. Behavioral
reactivity to stress: amplification of stress-induced noradrenergic activation elicits a galanin-mediated anxiolytic effect in central amygdala.
Pharmacol Biochem 71: 407–417, 2002.
126. Kirouac GJ, Ciriello J. Medullary inputs to nucleus accumbens neurons. Am J Physiol Regul Integr Comp Physiol 273: R2080 –R2088,
1997.
127. Kitazawa S, Shioda S, Nakai Y. Catecholaminergic innervation of
neurons containing corticotropin-releasing factor in the paraventricular
nucleus of the rat hypothalamus. Acta Anat (Basel) 129: 337–343, 1987.
128. Koob GF. Corticotropin-releasing factor, norepinephrine, and stress.
Biol Psychiatry 46: 1167–1180, 1999.
129. Kreek MJ, Koob GF. Drug dependence: stress and dysregulation of
brain reward pathways. Drug Alcohol Depend 51: 23–47, 1998.
130. Krukoff TL, Harris KH, Jhamandas JH. Efferent projections from the
parabrachial nucleus demonstrated with the anterograde tracer Phaseolus
vulgaris leucoagglutinin. Brain Res Bull 30: 163–172, 1993.
131. Laorden ML, Fuertes G, González-Cuello A, Milanés MV. Changes
in catecholaminergic pathways inervating paraventricular nucleus and
pituitary-adrenal axis responses during morphine dependence: implication of ␣1- and ␣2-adrenoceptors. J Pharmacol Exp 293: 578 –584, 2000.
132. Laorden ML, Núñez C, Almela P, Milanés MV. Morphine withdrawalinduced c-fos expression in the hypothalamic paraventricular nucleus is
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
96.
induced neuropeptide Y and agouti gene-related protein messenger ribonucleic acid expression in the arcuate nucleus. Endocrinology 144:
75–83, 2003.
Fuentealba JA, Forray MI, Gysling K. Chronic morphine treatment
and withdrawal increase extracellular levels of norepinephrine in the rat
bed nucleus of the stria terminalis. J Neurochem 75: 741–748, 2000.
Fuertes G, Laorden ML, Milanés MV. Noradrenergic and dopaminergic activity in the hypothalamic paraventricular nucleus after naloxoneinduced morphine withdrawal. Neuroendocrinology 71: 60 –67, 2000.
Füzesi T, Wittmann G, Lechan RM, Liposits Z, Fekete C. Noradrenergic innervation of hypophysiotropic thyrotropin-releasing hormone
synthesizing neurons in rats. Brain Res 1294: 38 –44, 2009.
Gabbott PL, Warner T, Busby SJ. Catecholaminergic neurons in
medullary nuclei are among the post-synaptic targets of descending
projections from infralimbic area 25 of the rat medial prefrontal cortex.
Neuroscience 144: 623–635, 2007.
Gaillet S, Alonso G, LeBorgne R, Barbanel G, Malaval F, Assenmacher I, Szafarczyk A. Effects of discrete lesions in the ventral noradrenergic ascending bundle on the corticotropic stress response depend on
the site of the lesion and on the plasma levels of adrenal steroids.
Neuroendocrinology 58: 408 –419, 1993.
Gaykema RP, Park SM, McKibbin CR, Goehler LE. Lipopolysaccharide suppresses activation of the tuberomammillary histaminergic
system concomitant with behavior: a novel target of immune-sensory
pathways. Neuroscience 152: 273–287, 2008.
Gaykema RPA, Chen CC, Goehler LE. Organization of immuneresponsive medullary projections to the bed nucleus of the stria terminalis, central amygdala, and paraventricular nucleus of the hypothalamus:
evidence for parallel viscerosensory pathways in the rat brain. Brain Res
1130: 130 –145, 2007.
Gaykema RPA, Daniels TE, Shapiro NJ, Thacker GC, Park SM,
Goehler LE. Immune challenge and satiety-related activation of both
distinct and overlapping neuronal populations in the brainstem indicate
parallel pathways for viscerosensory signaling. Brain Res 1294: 61–79,
2009.
Geerling JC, Kawata M, Loewy AD. Aldosterone-sensitive neurons in
the rat central nervous system. J Comp Neurol 494: 515–527, 2006.
Geerling JC, Shin JW, Chimenti PC, Loewy AD. Paraventricular
hypothalamic nucleus: axonal projections to the brainstem. J Comp
Neurol 518: 1460 –1499, 2010.
Härfstrand A, Fuxe K, Cintra A, Agnati LF, Zini I, Wikström AC,
Okret S, Yu ZY, Goldstein M, Steinbusch H, Verhofstad A, Gustafsson JA. Glucocorticoid receptor immunoreactivity in monoaminergic
neurons of rat brain. Proc Natl Acad Sci USA 83: 9779 –9783, 1986.
Harris GC, Aston-Jones G. Activation in extended amygdala corresponds to altered hedonic processing during protracted morphine withdrawal. Behav Brain Res 176: 251–258, 2007.
Harro J, Oreland L, Vasar E, Bradwejn J. Impaired exploratory
behavior after DSP-4 treatment in rats: implications for the increased
anxiety after noradrenergic denervation. Eur Neuropsychopharmacol 5:
447–455, 1995.
Herbert H, Moga MM, Saper CB. Connections of the parabrachial
nucleus with the nucleus of the solitary tract and the medullary reticular
formation in the rat. J Comp Neurol 293: 540 –580, 1990.
Herbert H, Saper CB. Cholecystokinin-, galanin-, and corticotropinreleasing factor-like immunoreactive projections from the nucleus of the
solitary tract to the parabrachial nucleus in the rat. J Comp Neurol 293:
581–598, 1990.
Herman JP, Cullinan WE. Neurocircuitry of stress: central control of
the hypothalamo-pituitary-adrenocortical axis. Trends Neurosci 20: 78 –
84, 1997.
Herman JP, Figueiredo H, Mueller NK, Ulrich-Lai Y, Ostrander
MM, Choi DC, Cullinan WE. Central mechanisms of stress integration:
hierarchical circuitry controlling hypothalamo-pituitary-adrenocortical
responsiveness. Front Neuroendocrinol 24: 151–180, 2003.
Herman JP, Ostrander MM, Mueller NK, Figueiredo H. Limbic
system mechanisms of stress regulation: hypothalamo-pituitary-adrenocortical axis. Prog Neuro-Psychopharm Biol Psych 29: 1201–1213,
2005.
Herman MA, Niedringhaus M, Alayan A, Verbalis JG, Sahibzada N,
Gillis RA. Characterization of noradrenergic transmission at the dorsal
motor nucleus of the vagus involved in reflex control of fundus tone. Am
J Physiol Regul Integr Comp Physiol 294: R720 –R729, 2008.
R231
Review
R232
133.
134.
135.
136.
137.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
dependent on the activation of catecholaminergic neurones. J Neurochem
83: 132–140, 2002.
Larsen PJ, Tang-Christensen M, Holst JJ, Orskov C. Distribution of
glucagon-like peptide-1 and other preproglucagon-derived peptides in rat
hypothalamus and brainstem. Neuroscience 77: 257–270, 1997.
Lawrence CB, Celsi F, Brennand J, Luckman SM. Alternative role for
prolactin-releasing peptide in the regulation of food intake. Nature 3:
645–646, 2000.
Leri F, Flores J, Rodaros D, Stewart J. Blockade of stress-induced but
not cocaine-induced reinstatement by infusion of noradrenergic antagonists into the bed nucleus of the stria terminalis or the central nucleus of
the amygdala. J Neurosci 22: 5713–5718, 2002.
Levin MC, Sawchenko PE, Howe PRC, Bloom SR, Polak JM.
Organization of galanin-immunoreactive inputs to the paraventricular
nucleus with special reference to their relationship to catecholaminergic
afferents. J Comp Neurol 261: 562–582, 1987.
Li HY, Ericsson A, Sawchenko PE. Distinct mechanisms underlie
activation of hypothalamic neurosecretory neurons and their medullary
catecholaminergic afferents in categorically different stress paradigms.
Proc Natl Acad Sci USA 93: 2359 –2364, 1996.
Li HY, Sawchenko PE. Hypothalamic effector neurons and extended
circuitries activated in “neurogenic” stress: a comparison of footshock
effects exerted acutely, chronically, and in animals with controlled
glucocorticoid levels. J Comp Neurol 393: 244 –266, 1998.
Liposits Z, Phelix C, Paull WK. Adrenergic innervation of corticotropin
releasing factor (CRF)-synthesizing neurons in the hypothalamic paraventricular nucleus of the rat. Histochemistry 84: 201–205, 1986.
Liubashina O, Jolkkonen E, Pitkanen A. Projections from the central
nucleus of the amygdala to the gastric related area of the dorsal vagal
complex: a Phaseolus vulgaris-leucoagglutinin study in rat. Neurosci
Lett 291: 85–88, 2000.
Loewy AD. Central autonomic pathways. In: Central Regulation of
Autonomic Functions, edited by Loewy AD and Spyer KM. New York:
Oxford University Press, 1990, p. 88 –103.
Lorden J, Oltmans GA, Margules DL. Central noradrenergic neurons:
differential effects on body weight of electrolytic and 6-hydroxydopamine lesions in rats. J Comp Physiol Psychol 90: 144 –155, 1976.
Lundy RF Jr, Norgren R. Gustatory system. In: The Rat Nervous
System (3rd ed.), edited by Paxinos G. San Diego, CA: Elsevier Academic, 2004, p. 890 –921.
Lynn RB, Kreider MS, Miselis RR. Thyrotropin-releasing hormoneimmunoreactive projections to the dorsal motor nucleus and the nucleus
of the solitary tract of the rat. J Comp Neurol 311: 271–288, 1991.
Maldonado R. Participation of noradrenergic pathways in the expression
of opiate withdrawal: biochemical and pharmacological evidence. Neurosci Biobehav Rev 21: 91–104, 1997.
Martinez-Peña-y-Valenzuela I, Rogers RC, Hermann GE, Travagli
RA. Norepinephrine effects on identified neurons of the rat dorsal motor
nucleus of the vagus. Am J Physiol Gastrointest Liver Physiol 286:
G333–G339, 2004.
Maruyama M, Matsumoto H, Fujiwara K, Noguchi J, Kitada C,
Fujino M, Inoue K. Prolactin-releasing peptide as a novel stress mediator in the central nervous system. Endocrinology 142: 2032–2038, 2001.
Mason ST, Fibiger HC. Current concepts. I. Anxiety: the locus coeruleus disconnection. Life Sci 25: 2141–2147, 1979.
Mayer EA, Saper CB (Editors). The Biological Basis for Mind Body
Interactions. Amsterdam: Elsevier, 2000.
Mehendale S, Xie JT, Aung HH, Guan XF, Yuan CS. Nucleus
accumbens receives gastric vagal inputs. Acta Pharmacol Sin 25: 271–
275, 2004.
Mejías-Aponte CA, Drouin C, Astron-Jones G. Adrenergic and noradrenergic innervation of the midbrain ventral tegmental area and retrorubal field: prominent inputs from medullary homeostatic centers. J
Neurosci 29: 3613–3626, 2009.
Menétrey D, Basbaum AI. Spinal and trigeminal projections to the
nucleus of the solitary tract: a possible substrate for somatovisceral and
viscerovisceral reflex activation. J Comp Neurol 255: 439 –450, 1987.
Menétrey D, dePommery J. Origins of spinal ascending pathways that
reach central areas involved in visceroception and visceronociception in
the rat. Eur J Neurosci 3: 249 –259, 1991.
Meyer H, Palchaudhuri M, Scheinin M, Flügge G. Regulation of
␣2A-adrenoceptor expression by chronic stress in neurons of the brain
stem. Brain Res 880: 147–158, 2000.
AJP-Regul Integr Comp Physiol • VOL
155. Michaloudi HC, Majdoubi ME, Poulain DA, Papadopoulos GC,
Theodosis DT. The noradrenergic innervation of identified hypothalamic
magnocellular somata and its contribution to lactation-induced synaptic
plasticity. J Neuroendocrinol 9: 17–23, 1997.
156. Milan MJ, Millan MH, Herz A. The role of the ventral noradrenergic
bundle in relation to endorphins in the control of core temperature,
open-field and ingestive behaviour in the rat. Brain Res 263: 283–294,
1983.
157. Millhorn DE, Seroogy K, Hökfelt T, Schmued LC, Terenius L,
Buchan A, Brown JC. Neurons of the ventral medulla oblongata that
contain both somatostatin and enkephalin immunoreactivities project to
nucleus tractus solitarii and spinal cord. Brain Res 424: 99 –108, 1987.
158. Milner TA, Joh TH, Pickel VM. Tyrosine hydroxylase in the rat
parabrachial region: ultrastructural localization and extrinsic sources of
immunoreactivity. J Neurosci 6: 2585–2603, 1986.
159. Miranda MI, Ferreira G, Ramirez-Lugo L, Bermúdez-Rattoni F.
Glutamatergic activity in the amygdala signals visceral input during taste
memory formation. Proc Natl Acad Sci USA 99: 11417–11422, 2002.
160. Miyahara S, Oomura Y. Inhibitory action of the ventral noradrenergic
bundle on the lateral hypothalamic neurons through ␣-noradrenergic
mechanisms in the rat. Brain Res 234: 459 –463, 1982.
161. Miyashita T, Williams CL. Enhancement of noradrenergic neurotransmission in the nucleus of the solitary tract modulates memory storage
processes. Brain Res 987: 164 –175, 2003.
162. Miyashita T, Williams CL. Glutamatergic transmission in the nucleus
of the solitary tract modulates memory through influences on amygdala
noradrenergic systems. Behav Neurosci 116: 13–21, 2002.
163. Moore RY, Card JP. Noradrenaline-containing neuron systems. In:
Handbook of Chemical Neuroanatomy, edited by Bj¨orklund A and
Ḧokfelt T. Amsterdam: Elsevier, 1984, p. 123–156.
164. Morilak DA, Barrera G, Echevarria DJ, Garcia AS, Hernandez A,
Ma S, Petre CO. Role of brain norepinephrine in the behavioral response
to stress. Prog Neuro-Psychopharm Biol Psych 29: 1214 –1224, 2005.
165. Mounier F, Bluet-Pajot MT, Durand D, Kordon C, Epelbaum J.
␣1-Noradrenergic inhibition of growth hormone secretion is mediated
through the paraventricular hypothalamic nucleus in male rats. Neuroendocrinology 59: 29 –34, 1994.
166. Mtui EP, Anwar M, Reis DJ, Ruggiero DA. Medullary visceral reflex
circuits: local afferents to nucleus tractus solitarii synthesize catecholamines and project to thoracic spinal cord. J Comp Neurol 351:
5–26, 1995.
167. Myers EA, Banihashemi L, Rinaman L. The anxiogenic drug yohimbine activates central viscerosensory circuits in rats. J Comp Neurol 492:
426 –441, 2005.
168. Myers EA, Rinaman L. Trimethylthiazoline supports conditioned flavor
avoidance and activates viscerosensory, hypothalamic, and limbic circuits in rats. Am J Physiol Regul Integr Comp Physiol 288: R1716 –
R1726, 2005.
169. Myers EA, Rinaman L. Viscerosensory activation of noradrenergic
inputs to the amygdala in rats. Physiol Behav 77: 723–729, 2002.
170. Myers RD, McCaleb ML. Feeding: satiety signals from intestine trigger
brain’s noradrenergic mechanisms. Science 209: 1035–1037, 1980.
171. Navarro-Zaragoza J, Núñez C, Laorden ML, Milanés MV. Effects of
corticotropin-releasing factor receptor-1 antagonists on the brain stress
system responses to morphine withdrawal. Mol Pharmacol 77: 864 –873,
2010.
172. Núñez C, Földes A, Pérez-Flores D, García-Borrón JC, Laorden ML,
Kovács KJ, Milanés MV. Elevated glucocorticoid levels are responsible
for induction of tyrosine hydroxylase mRNA expression, phosphorylation, and enzyme activity in the nucleus of the solitary tract during
morphine withdrawal. Endocrinology 150: 3118 –3127, 2009.
173. Núñez C, Martín F, AAF, Laorden ML, Kovács KJ, Milanés MV.
Induction of FosB/DeltaFosB in the brain stress system-related structures
during morphine dependence and withdrawal. J Neurochem 114: 475–
487, 2010.
174. Oh-IS, Shimizu H, Satoh T, Okada S, Adachi S, Inoue K, Eguchi H,
Yamamoto M, Imaki T, Hashimoto K, Tsuchiya T, Monden T,
Horiguchi K, Yamada M, Mori M. Identification of nesfatin-1 as a
satiety molecule in the hypothalamus. Nature (London) 443: 709 –712,
2006.
175. Olschowka JA, Molliver ME, Grzanna R, Rice FL, Coyle JJ. Ultrastructural demonstration of noradrenergic synapses in the rat central
neuron system by dopamine-␤-hydroxylase immunocytochemistry. J
Histochem Cytochem 29: 271–289, 1981.
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
138.
HINDBRAIN NORADRENERGIC A2 NEURONS
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
AJP-Regul Integr Comp Physiol • VOL
199. Ricardo JA, Koh ET. Anatomical evidence of direct projections from
the nucleus of the solitary tract to the hypothalamus, amygdala, and other
forebrain structures in the rat. Brain Res 153: 1–26, 1978.
200. Riche D, Pommery JD, Menetrey D. Neuropeptides and catecholamines in efferent projections of the nuclei of the solitary tract in
the rat. J Comp Neurol 293: 399 –424, 1990.
201. Rinaman L. Ascending projections from the caudal visceral nucleus of
the solitary tract to brain regions involved in food intake and energy
expenditure. Brain Res 1350: 18 –34, 2010.
202. Rinaman L. Hindbrain contributions to anorexia. Am J Physiol Regul
Integr Comp Physiol 287: R1035–R1036, 2004.
203. Rinaman L. Hindbrain noradrenergic lesions attenuate anorexia and
alter central cFos expression in rats after gastric viscerosensory stimulation. J Neurosci 23: 10084 –10092, 2003.
204. Rinaman L. Oxytocinergic inputs to the nucleus of the solitary tract and
dorsal motor nucleus of the vagus in neonatal rats. J Comp Neurol 399:
101–109, 1998.
205. Rinaman L. Postnatal development of hypothalamic inputs to the dorsal
vagal complex in rats. Physiol Behav 79: 65–70, 2003.
206. Rinaman L. Visceral sensory inputs to the endocrine hypothalamus.
Front Neuroendocrinol 28: 50 –60, 2007.
207. Rinaman L, Baker EA, Hoffman GE, Stricker EM, Verbalis JG.
Medullary c-Fos activation in rats after ingestion of a satiating meal. Am
J Physiol Regul Integr Comp Physiol 275: R262–R268, 1998.
208. Rinaman L, Card JP, Schwaber JS, Miselis RR. Ultrastructural
demonstration of a gastric monosynaptic vagal circuit in the nucleus of
the solitary tract in rat. J Neurosci 9: 1985–1996, 1989.
209. Rinaman L, Dzmura V. Experimental dissociation of neural circuits
underlying conditioned avoidance and hypophagic responses to lithium
chloride. Am J Physiol Regul Integr Comp Physiol 293: R1495–R1503,
2007.
210. Rinaman L, Hoffman GE, Dohanics J, Le WW, Stricker EM, Verbalis JG. Cholecystokinin activates catecholaminergic neurons in the
caudal medulla that innervate the paraventricular nucleus of the hypothalamus in rats. J Comp Neurol 360: 246 –256, 1995.
211. Rinaman L, Schwartz GJ. Anterograde transneuronal viral tracing of
central viscerosensory pathways in rats. J Neurosci 24: 2782–2786, 2004.
212. Rinaman L, Stricker EM, Hoffman GE, Verbalis JG. Central c-fos
expression in neonatal and adult rats after subcutaneous injection of
hypertonic saline. Neuroscience 79: 1165–1175, 1997.
213. Rinaman L, Verbalis JG, Stricker EM, Hoffman GE. Distribution and
neurochemical phenotypes of caudal medullary neurons activated to
express cFos following peripheral administration of cholecystokinin. J
Comp Neurol 338: 475–490, 1993.
214. Ritter S, Bugarith K, Dinh TT. Immunotoxic destruction of distinct
catecholamine subgroups produces selective impairment of glucoregulatory responses and neuronal activation. J Comp Neurol 432: 197–216,
2001.
215. Ritter S, Watts AG, Dinh TT, Sanchez-Watts G, Pedrow C. Immunotoxin lesion of hypothalamically projecting norepinephrine and epinephrine neurons differentially affects circadian and stressor-stimulated
corticosterone secretion. Endocrinology 144: 1357–1367, 2003.
216. Ritter S, Wise D, Stein L. Neurochemical regulation of feeding in the
rat: facilitation by ␣-noradrenergic, but not dopaminergic, receptor stimulants. J Comp Physiol Psychol 88: 778 –784, 1975.
217. Rivier CL, Plotsky PM. Mediation by corticotropin-releasing factor of
adenohypophysial hormone secretion. Annu Rev Physiol 48: 475–494,
1986.
218. Rogers RC, Travagli RA, Hermann GE. Noradrenergic neurons in the
rat solitary nucleus participate in the esophageal-gastric relaxation reflex.
Am J Physiol Regul Integr Comp Physiol 285: R479 –R489, 2003.
219. Roozendaal B, Williams CL, McGaugh JL. Glucocorticoid receptor
activation in the rat nucleus of the solitary tract facilitates memory
consolidation: involvement of the basolateral amygdala. Eur J Neurosci
11: 1317–1323, 1999.
220. Rosin DL, Chang DA, Guyenet PG. Afferent and efferent connections
of the rat retrotrapezoid nucleus. J Comp Neurol 499: 64 –89, 2006.
221. Rosin DL, Zeng D, Stornetta RL, Norton FR, Riley T, Okusa MD,
Guyenet PG, Lynch KR. Immunohistochemical localization of ␣2aadrenergic receptors in catecholaminergic and other brainstem neurons in
the rat. Neuroscience 56: 139 –155, 1993.
222. Saha S, Batten TFC, Henderson Z. A GABAergic projection from the
central nucleus of the amygdala to the nucleus of the solitary tract: a
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
176. Olson VG, Heusner CL, Bland RJ, During MJ, Weinshenker D,
Palmiter RD. Role of noradrenergic signaling by the nucleus tractus
solitarius in mediating opiate reward. Science 311: 1017–1020, 2006.
177. Onaka T. Neural pathways controlling central and peripheral oxytocin
release during stress. J Neuroendocrinol 16: 308 –312, 2004.
178. Onaka T, Palmer JR, Yagi K. A selective role of brainstem noradrenergic neurons in oxytocin release from the neurohypophysis following
noxious stimuli in the rat. Neurosci Res 25: 67–75, 1996.
179. Onaka T, Yagi K. Role of noradrenergic projections to the bed nucleus
of the stria terminalis in neuroendocrine and behavioral responses to
fear-related stimuli in rats. Brain Res 788: 287–293, 1998.
180. Otake K, Reis DJ, Ruggiero DA. Afferents to the midline thalamus
issue collaterals to the nucleus tractus solitarii: an anatomical basis for
thalamic and visceral reflex integration. J Neurosci 14: 5694 –5707,
1994.
181. Pacak K, Palkovits M, Kopin IJ, Goldstein DS. Stress-induced norepinephrine release in the hypothalamic paraventricular nucleus and
pituitary-adrenocortical and sympathoadrenal activity: in vivo microdialysis studies. Front Neuroendocrinol 16: 89 –150, 1995.
182. Panula P, Pirvola U, Auvinen S, Airaksinen MS. Histamine-immunoreactive nerve fibers in the rat brain. Neuroscience 28: 585–610, 1989.
183. Pardon MC, Ma S, Morilak DA. Chronic cold stress sensitizes brain
noradrenergic reactivity and noradrenergic facilitation of the HPA stress
response in Wistar Kyoto rats. Brain Res 971: 55–65, 2003.
184. Pearson RJ, Gatti PJ, Sahibzada N, Massari VJ, Gillis RA. Ultrastructural evidence for selective noradrenergic innervation of CNS vagal
projections to the fundus of the rat. Auton Neurosci 136: 31–42, 2007.
185. Petrov T, Krukoff TL, Jhamandas JH. Branching projections of
catecholaminergic brainstem neurons to the paraventricular hypothalamic nucleus and the central nucleus of the amygdala in the rat. Brain
Res 609: 81–92, 1993.
186. Phelix CF, Liposits Z, Paull WK. Catecholamine-CRF synaptic interaction in a septal bed nucleus: afferents of neurons in the bed nucleus of
the stria terminalis. Brain Res Bull 33: 109 –119, 1994.
187. Phelix CF, Liposits Z, Paull WK. Monoamine innervation of bed
nucleus of stria terminalis: an electron microscopic investigation. Brain
Res Bull 28: 949 –965, 1992.
188. Phelix CF, Paull WK. Demonstration of distinct corticotropin releasing
factor-containing neuron populations in the bed nucleus of the stria
terminalis. A light and electron microscopic immunocytochemical study
in the rat. Histochemistry 94: 345–364, 1990.
189. Pickel VM, Bockstaele EJv Chan J, Cestari DM. Amygdala efferents
form inhibitory-type synapses with a subpopulation of catecholaminergic
neurons in the rat nucleus tractus solitarius. J Comp Neurol 362: 510 –
523, 1995.
190. Pickel VM, Nirenberg MJ, Milneer TA. Ultrastructural view of central
catecholamine transmission: immunocytochemical localization of synthesizing enzymes, transporters and receptors. J Neurocytol 25: 843–856,
1996.
191. Plotsky P. Facilitation of immunoreactive corticotropin-releasing factor
secretion into the hypophysial-portal circulation after activation of catecholaminergic pathways or central norepinephrine injection. Endocrinology 121: 924 –930, 1987.
192. Plotsky PM, Cunningham ET Jr, Widmaier EP. Catecholaminergic
modulation of corticotropin-releasing factor and adrenocorticotropin secretion. Endocr Rev 10: 437–458, 1989.
193. Price JL. Free will versus survival: brain systems that underlie intrinsic
constraints on behavior. J Comp Neurol 493: 132–139, 2005.
194. Prinz JJ. Gut Reactions: A Perceptual Theory of Emotion. New York
Oxford University Press, 2004.
195. Raby WN, Renaud LP. Dorsomedial medulla stimulation activates rat
supraoptic oxytocin and vasopressin neurones through different pathways. J Physiol 417: 279 –294, 1989.
196. Raby WN, Renaud LP. Nucleus tractus solitarius innervation of supraoptic nucleus: anatomical and electrophysiological studies in the rat
suggest differential innervation of oxytocin and vasopressin neurons.
Prog Brain Res 81: 319 –327, 1989.
197. Reyes BAS, Bockstaele EJV. Divergent projections of catecholaminergic neurons in the nucleus of the solitary tract to limbic forebrain and
medullary autonomic brain regions. Brain Res 1117: 69 –79, 2006.
198. Reyes BAS, Tsukamura H, I’Anson H, Amelita M, Estacio C,
Hirunagi K, Maeda KI. Temporal expression of estrogen receptor-␣ in
the hypothalamus and medulla oblongata during fasting: a role of
noradrenergic neurons. J Endocrinol 190: 593–600, 2006.
R233
Review
R234
223.
224.
225.
226.
228.
229.
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
240.
241.
242.
243.
244.
combined anterograde tracing and electron microscopic immunohistochemical study. Neuroscience 99: 613–626, 2000.
Sahuque LL, Kullberg EF, Mcgeehan AJ, Kinder JR, Hicks MP,
Blanton MG, Janak PH, Olive MF. Anxiogenic and aversive effects of
corticotropin-releasing factor (CRF) in the bed nucleus of the stria
terminalis in the rat: role of CRF receptor subtypes. Psychopharmacology (Berl) 186: 122–132, 2006.
Sartor DM, Verberne AJM. The sympathoinhibitory effects of systemic cholecystokinin are dependent on neurons in the caudal ventrolateral medulla in the rat. Am J Physiol Regul Integr Comp Physiol 291:
R1390 –R1398, 2006.
Sawchenko PE. Central connections of the sensory and motor nuclei of
the vagus nerve. J Auton Nerv Syst 9: 13–26, 1983.
Sawchenko PE, Arias C, Bittencourt JC. Inhibin B, somatostatin, and
enkephalin immunoreactivities coexist in caudal medullary neurons that
project to the paraventricular nucleus of the hypothalamus. J Comp
Neurol 291: 269 –280, 1990.
Sawchenko PE, Benoit R, Brown MR. Somatostatin 28-immunoreactive inputs to the paraventricular and supraoptic nuclei: principal origin
from non-aminergic neurons in the nucleus of the solitary tract. J Chem
Neuroanat 1: 81–94, 1988.
Sawchenko PE, Brown ER, Chan RKW, Ericsson A, Li HY, Roland
BL, Kovacs KJ. The paraventricular nucleus of the hypothalamus and
the functional neuroanatomy of visceromotor responses to stress. Prog
Brain Res 107: 201–222, 1996.
Sawchenko PE, Li HY, Ericsson A. Circuits and mechanisms governing hypothalamic responses to stress: a tale of two paradigms. In: Prog
Brain Res, edited by Mayer EA and Saper CB. Amsterdam: Elsevier
Science, 2000, p. 61–78.
Sawchenko PE, Swanson LW. Central noradrenergic pathways for the
integration of hypothalamic neuroendocrine and autonomic responses.
Science 214: 685–687, 1981.
Sawchenko PE, Swanson LW. The organization of forebrain afferents
to the paraventricular and supraoptic nuclei of the rat. J Comp Neurol
218: 121–144, 1983.
Sawchenko PE, Swanson LW. The organization of noradrenergic pathways from the brainstem to the paraventricular and supraoptic nuclei in
the rat. Brain Res Rev 4: 275–325, 1982.
Sawchenko PE, Swanson LW, Grzanna R, Howe PRC, Bloom SR,
Polak JM. Colocalization of neuropeptide Y immunoreactivity in brainstem catecholaminergic neurons that project to the paraventricular nucleus of the hypothalamus. J Comp Neurol 241: 138 –153, 1985.
Schiltz JC, Sawchenko PE. Specificity and generality of the involvement of catecholaminergic afferents in hypothalamic responses to immune insults. J Comp Neurol 502: 455–467, 2007.
Serova LI, Harris HA, Maharjan S, Sabban EL. Modulation of
responses to stress by estradiol benzoate and selective estrogen receptor
agonists. J Endocrinol 205: 253–262, 2010.
Shaham Y, Erb S, Stewart J. Stress-induced relapse to heroin and
cocaine seeking in rats: a review. Brain Res Rev 33: 13–33, 2000.
Shaham Y, Highfield D, Delfs J, Leung S, Stewart J. Clonidine blocks
stress-induced reinstatement of heroin seeking in rats: an effect independent of locus coeruleus noradrenergic neurons. Eur J Neurosci 12:
292–302, 2000.
Shapiro RE, Miselis RR. The central neural connections of the area
postrema of the rat. J Comp Neurol 234: 344 –364, 1985.
Shapiro RE, Miselis RR. The central organization of the vagus nerve
innervating the stomach of the rat. J Comp Neurol 238: 473–488, 1985.
Shioda S, Nakai Y. Medullary synaptic inputs to thyrotropin-releasing
hormone (TRH)-containing neurons in the hypothalamus: an ultrastructural study combining WGA-HRP anterograde tracing with TRH immunocytochemistry. Brain Res 625: 9 –15, 1993.
Siever L, Insel T, Uhde T. Noradrenergic challenges in the affective
disorders. J Clin Psychopharmacol 1: 193–206, 1981.
Siever LJ, Uhde TW, Silberman EK, Jimerson DC, Aloi JA, Post
RM, Murphy DL. The growth hormone response to clonidine as a probe
of noradrenergic receptor responsiveness in affective disorder patients
and controls. Psychiatr Res 6: 171–183, 1982.
Simonian SX, Delaleu B, Caraty A, Herbison AE. Estrogen receptor
expression in brainstem noradrenergic neurons of the sheep. Neuroendocrinology 67: 392–402, 1998.
Smith RJ, Aston-Jones G. Noradrenergic transmission in the extended
amygdala: role in increased drug-seeking and relapse during protracted
drug abstinence. Brain Struct Funct 213: 43–61, 2008.
AJP-Regul Integr Comp Physiol • VOL
245. Sofroniew MV, Schrell U. Evidence for a direct projection from oxytocin and vasopressin neurons in the hypothalamic paraventricular nucleus to the medulla oblongata: immunohistochemical visualization of
both the horseradish peroxidase transported and the peptide produced by
the same neurons. Neurosci Lett 22: 211–217, 1981.
246. Spyer KM. The central nervous organization of reflex circulatory control. In: Central Regulation of Autonomic Functions, edited by Loewy
AD and Spyer KM. New York: Oxford University Press, 1990, p.
168 –188.
247. Stone EA, Quartermain D, Lin Y, Lehmann ML. Central ␣1-adrenergic system in behavioral activity and depression. Biochem Pharmacol
73: 1063–1075, 2007.
248. Stornetta RL, Sevigny CP, Guyenet PG. Vesicular glutamate transporter DNPI/VGLUT2 mRNA is present in C1 and several other groups
of brainstem catecholaminergic neurons. J Comp Neurol 444: 191–206,
2002.
249. Swanson LW, Petrovich GD. What is the amygdala? Trends Neurosci
21: 323–331, 1998.
250. Szafarczyk A, Alonso G, Ixart G, Malaval F, Assenmacher I. Diurnalstimulated and stress-induced ACTH release in rats is mediated by
ventral noradrenergic bundle. Am J Physiol Endocrinol Metab 249:
E219 –E226, 1985.
251. Taché Y, Garrick T, Raybould H. Central nervous system action of
peptides to influence gastrointestinal motor function. Gastroenterology
98: 517–528, 1990.
252. Terenzi MG, Ingram CD. A combined immunocytochemical and retrograde tracing study of noradrenergic connections between the caudal
medulla and bed nucleus of the stria terminalis. Brain Res 672: 289 –297,
1995.
253. Ter Horst GJ, Boer PD, Luiten PGM, Willigen JDV. Ascending
projections from the solitary tract nucleus to the hypothalamus. A
Phaseolus vulgaris lectin tracing study in the rat. Neuroscience 31:
785–797, 1989.
254. Ter Horst GJ, Streefland C. Ascending projections of the solitary tract
nucleus. In: Nucleus of the Solitary Tract, edited by Barraco IRA. Boca
Raton, FL: CRC, 1994, p. 93–103.
255. Thayer JF, Lane RD. A model of neurovisceral integration in emotional
regulation and dysregulation. J Affect Disord 61: 201–216, 2000.
256. Thor KB, Helke CJ. Serotonin and substance P colocalization in
medullary projections to the nucleus tractus solitarius: dual-colour immunohistochemistry combined with retrograde tracing. J Chem Neuroanat 2: 139 –148, 1989.
257. Toufexis DJ, Thrivikraman KV, Plotsky PM, Morilak DA, Huang N,
Walker CD. Reduced noradrenergic tone to the hypothalamic paraventricular nucleus contributes to the stress hyporesponsiveness of lactation.
J Neuroendocrinol 10: 417–427, 1998.
258. Toufexis DJ, Walker CD. Noradrenergic facilitation of the adrenocorticotropin response to stress is absent during lactation in the rat. Brain
Res 737: 71–77, 1996.
259. Travagli RA, Hermann GE, Browning KN, Rogers RC. Brainstem
circuits regulating gastric function. Annu Rev Physiol 68: 279 –305, 2006.
260. Tucker DC, Saper CB, Ruggiero DA, Reis DJ. Organization of central
adrenergic pathways: I. Relationships of ventrolateral medullary projections to the hypothalamus and spinal cord. J Comp Neurol 259: 591–603,
1987.
261. Uchida K, Kobayashi D, Das G, Onaka T, Inoue K, Itoi K. Participation of the prolactin-releasing peptide-containing neurones in caudal
medulla in conveying haemorrhagic stress-induced signals to the paraventricular nucleus of the hypothalamus. J Neuroendocrinol 22: 33–42,
2010.
262. Uchoa ET, Sabino HA, Ruginsk SG, Antunes-Rodrigues J, Elias LL.
Hypophagia induced by glucocorticoid deficiency is associated with an
increased activation of satiety-related responses. J Appl Physiol 106:
596 –604, 2009.
263. Ueta Y, Kannan H, Higuchi T, Negoro H, Yamashita H. CCK-8
excites oxytocin-secreting neurons in the paraventricular nucleus in
rats-possible involvement of noradrenergic pathway. Brain Res Bull 32:
453–459, 1993.
264. Van Bockstaele EJ, Bajic D, Proudfit H, Valentino RJ. Topographic
architecture of stress-related pathways targeting the noradrenergic locus
coeruleus. Physiol Behav 73: 273–283, 2001.
265. Van Bockstaele EJ, Peoples J, Telegan P. Efferent projections of the
nucleus of the solitary tract to peri-locus coeruleus dendrites in rat brain:
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
227.
HINDBRAIN NORADRENERGIC A2 NEURONS
Review
HINDBRAIN NORADRENERGIC A2 NEURONS
266.
267.
268.
269.
270.
272.
273.
274.
275.
276.
277.
278.
279.
280.
AJP-Regul Integr Comp Physiol • VOL
281. Williams CL, Men D, Clayton EC, Gold PE. Norepinephrine release in
the amygdala after systemic injection of epinephrine or escapable footshock: contribution of the nucleus of the solitary tract. Behav Neurosci
112: 1414 –1422, 1998.
282. Willoughby JO, Chapman IM, Kapoor R. Local hypothalamic adrenoceptor activation in rat: ␣1 inhibits and ␣2 stimulates growth hormone
secretion. Neuroendocrinology 57: 687–692, 1993.
283. Wittman G. Regulation of hypophysiotropic corticotrophin-releasing
hormone- and thyrotrophin-releasing hormone-synthesising neurones by
brainstem catecholaminergic neurones. J Neuroendocrinol 20: 952–960,
2008.
284. Yamada T, Mochiduki A, Sugimoto Y, Suzuki Y, Itoi K, Inoue K.
Prolactin-releasing peptide regulates the cardiovascular system via corticotrophin-releasing hormone. J Neuroendocrinol 21: 586 –593, 2009.
285. Yamano M, Bai F, Tohyam M, Shiotani Y. Ultrastructural evidence of
direct synaptic contact of catecholamine terminals with oxytocin-containing neurons in the parvocellular portion of the rat hypothalamic
paraventricular nucleus. Brain Res 336: 176 –179, 1985.
286. Yamashita H, Kannan H, Ueta Y. Involvement of caudal ventrolateral
medulla neurons in mediating visceroreceptive information to the hypothalamic paraventricular nucleus. Prog Brain Res 81: 293–302, 1989.
287. Yang L, Scott KA, Hyun J, Tamashiro KL, Tray N, Moran TH, Bi
S. Role of dorsomedial hypothalamic neuropeptide Y in modulating food
intake and energy balance. J Neurosci 29: 179 –190, 2009.
288. Yang SN, Bunnemann B, Cintra A, Fuxe K. Localization of neuropeptide Y Y1 receptor-like immunoreactivity in catecholaminergic neurons
of the rat medulla oblongata. Neuroscience 73: 519 –530, 1996.
289. Young EA, Abelson JL, Cameron OG. Interaction of brain noradrenergic system and the hypothalamic-pituitary-adrenal (HPA) axis in man.
Psychoneuroendocrinology 30: 807–814, 2005.
290. Yu G, Sharp B. Nicotine self-administration diminishes stress-induced
norepinephrine secretion but augments adrenergic-responsiveness in the
hypothalamic paraventricular nucles and enhances adrenocorticotropic
hormone and corticosterone release. J Neurochem 112: 1327–1237,
2010.
291. Zagon A, Rocha I, Ishizuka K, Spyer KM. Vagal modulation of
responses elicited by stimulation of the aortic depressor nerve in neurons
of the rostral ventrolateral medulla oblongata in the rat. Neuroscience 92:
889 –899, 1999.
292. Zhang JF, Zheng F. The role of paraventricular nucleus of hypothalamus in stress-ulcer formation in rats. Brain Res 761: 203–209, 1997.
293. Zhao R, Chen H, Sharp BM. Nicotine-induced norepinephrine release
in hypothalamic paraventricular nucleus and amygdala is mediated by
N-methyl-D-aspartate receptors and nitric oxide in the nucleus tractus
solitarius. J Pharmacol Exp 320: 837–844, 2007.
294. Zheng H, Patterson LM, Berthoud HR. Orexin-A projections to the
caudal medulla and orexin-induced c-fos expression, food intake, and
autonomic function. J Comp Neurol 485: 127–142, 2005.
295. Zheng H, Patterson LM, Rhodes CJ, Louis GW, Skibicka KP, Grill
HJ, Myers MG Jr, Berthoud HR. A potential role for hypothalamomedullary POMC projections in leptin-induced suppression of food
intake. Am J Physiol Regul Integr Comp Physiol 298: R720 –R728, 2010.
300 • FEBRUARY 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.33.1 on June 11, 2017
271.
evidence for a monosynaptic pathway. J Comp Neurol 412: 410 –428,
1999.
Van der Kooy D, Koda LK, McGinty JF, Gerfen CR, Bloom FE. The
organization of projections from the cortex, amygdala, and hypothalamus
to the nucleus of the solitary tract in rat. J Comp Neurol 1984: 1–24,
1984.
Verbalis JG, Hoffman GE, Sherman TG. Use of immediate early
genes as markers of oxytocin and vasopressin neuronal activation. Curr
Opin Endocrinol Diab 2: 157–168, 1995.
Verbalis JG, McHale CM, Gardiner TW, Stricker EM. Oxytocin and
vasopressin secretion in response to stimuli producing learned taste
aversions in rats. Behav Neurosci 100: 466 –475, 1986.
Verbalis JG, Stricker EM, Robinson AG, Hoffman GE. Cholecystokinin activates cFos expression in hypothalamic oxytocin and corticotropin-releasing hormone neurons. J Neuroendocrinol 3: 205–213, 1991.
Walker DL, Davis M. Double dissociation between the involvement of
the bed nucleus of the stria terminalis and the central nucleus of the
amygdala in startle increases produced by conditioned versus unconditioned fear. J Neurosci 17: 9375–9383, 1997.
Walker DL, Toufexis DJ, Davis M. Role of the bed nucleus of the stria
terminalis versus the amygdala in fear, stress, and anxiety. Eur J
Pharmacol 463: 199 –216, 2003.
Wang ZJ, Rao ZR, Shi JW. Tyrosein hydroxylase-, neurotensin-, or
cholecystokinin-containing neurons in the nucleus tractus solitarii send
projection fibers to the nucleus accumbens in the rat. Brain Res 578:
347–350, 1992.
Watanabe T, Nakagawa T, Yamamoto R, Maeda A, Minami M,
Satoh M. Involvement of noradrenergic system within the central nucleus of the amygdala in naloxone-precipitated morphine withdrawalinduced conditioned place aversion in rats. Psychopharmacology (Berl)
170: 80 –88, 2003.
Watts AG. The impact of physiological stimuli on the expression of
corticotropin-releasing hormone (CRH) and other neuropeptide genes.
Front Neuroendocrinol 17: 281–326, 1996.
Watts AG. Neuropeptides and the integration of motor responses to
dehydration. Annu Rev Neurosci 24: 357–384, 2001.
Weidenfeld J, Itzik A, Feldman S. Effects of glucocorticoids on the
adrenocortical axis responses to electrical stimulation of the amygdala
and the ventral noradrenergic bundle. Brain Res 754: 187–194, 1997.
Wellman PJ. Norepinephrine and the control of food intake. Nutrition
16: 837–842, 2000.
Wellman PJ, Davies BT. Suppression of feeding induced by phenylephrine microinjections within the paraventricular hypothalamus in rats.
Appetite 17: 121–128, 1991.
Williams CL, McGaugh JL. Reversible lesions of the nucleus of the
solitary tract attenuate the memory-modulating effects of posttraining
epinephrine. Behav Neurosci 107: 955–962, 1993.
Williams CL, Men D, Clayton EC. The effects of noradrenergic
activation of the nucleus tractus solitarius on memory and potentiating
norepinephrine release in the amygdala. Behav Neurosci 114: 1131–
1144, 2000.
R235